Next Article in Journal
Pomegranate Flower Extract—The Health-Promoting Properties Optimized by Application of the Box–Behnken Design
Next Article in Special Issue
Synthesis, Electrochemical and Fluorescence Properties of the First Fluorescent Member of the Ferrocifen Family and of Its Oxidized Derivatives
Previous Article in Journal
Production of Minor Ginsenosides from Panax notoginseng Flowers by Cladosporium xylophilum
Previous Article in Special Issue
Electrochemical Properties of Carbon Fibers from Felts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Properties of a Rhodium(III) Mono-Terpyridyl Complex and Use as a Catalyst for Light-Driven Hydrogen Evolution in Water

1
DCM, CNRS, Université Grenoble Alpes, 38000 Grenoble, France
2
SyMMES, IRIG, CEA, CNRS, Université Grenoble Alpes, 38000 Grenoble, France
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(19), 6614; https://doi.org/10.3390/molecules27196614
Submission received: 30 August 2022 / Revised: 23 September 2022 / Accepted: 30 September 2022 / Published: 5 October 2022

Abstract

:
Molecular hydrogen (H2) is considered one of the most promising fuels to decarbonize the industrial and transportation sectors, and its photocatalytic production from molecular catalysts is a research field that is still abounding. The search for new molecular catalysts for H2 production with simple and easily synthesized ligands is still ongoing, and the terpyridine ligand with its particular electronic and coordination properties, is a good candidate to design new catalysts meeting these requirements. Herein, we have isolated the new mono-terpyridyl rhodium complex, [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy), and shown that it can act as a catalyst for the light-induced proton reduction into H2 in water in the presence of the [Ru(bpy)3]Cl2 (Ru) photosensitizer and ascorbate as sacrificial electron donor. Under photocatalytic conditions, in acetate buffer at pH 4.5 with 0.1 M of ascorbate and 530 μM of Ru, the Rh-tpy catalyst produces H2 with turnover number versus catalyst (TONCat*) of 300 at a Rh concentration of 10 μM, and up to 1000 at a concentration of 1 μM. The photocatalytic performance of Ru/Rh-tpy/HA/H2A has been also compared with that obtained with the bis-dimethyl-bipyridyl complex [RhIII(dmbpy)2Cl2]+ (Rh2) as a catalyst in the same experimental conditions. The investigation of the electrochemical properties of Rh-tpy in DMF solvent reveals that the two-electrons reduced state of the complex, the square-planar [RhI(tpy)Cl] (RhI-tpy), is quantitatively electrogenerated by bulk electrolysis. This complex is stable for hours under an inert atmosphere owing to the π-acceptor property of the terpyridine ligand that stabilizes the low oxidation states of the rhodium, making this catalyst less prone to degrade during photocatalysis. The π-acceptor property of terpyridine also confers to the Rh-tpy catalyst a moderately negative reduction potential (Epc(RhIII/RhI) = −0.83 V vs. SCE in DMF), making possible its reduction by the reduced state of Ru, [RuII(bpy)(bpy•−)]+ (Ru) (E1/2(RuII/Ru) = −1.50 V vs. SCE) generated by a reductive quenching of the Ru excited state (*Ru) by ascorbate during photocatalysis. A Stern–Volmer plot and transient absorption spectroscopy confirmed that the first step of the photocatalytic process is the reductive quenching of *Ru by ascorbate. The resulting reduced Ru species (Ru) were then able to activate the RhIII-tpy H2-evolving catalyst by reduction generating RhI-tpy, which can react with a proton on a sub-nanosecond time scale to form a RhIII(H)-tpy hydride, the key intermediate for H2 evolution.

1. Introduction

Molecular hydrogen (H2) is considered as one of the best alternatives to fossil fuels to speed up the transition towards a carbon-neutral future [1]. However, 96% of hydrogen is produced from fossil fuels via a variety of processes, such as the steam reforming of methane, generating carbon dioxide at the same time. Therefore, to reach carbon neutrality, hydrogen should be cleanly produced using a renewable energy, such as sunlight, with water as a proton source. The pioneer work of Lehn and Sauvage demonstrated in the late 1970s that green hydrogen could be produced by the photo-induced reduction of protons via homogeneous photocatalytic systems [2,3]. These systems are usually composed of a sacrificial electron donor (SD) providing electrons to the system, a H2-evolving catalyst (Cat), and a photosensitizer (PS) promoting the photo-induced electron transfers between the three compounds. For forty years, many photocatalytic systems have been investigated and numerous molecular catalysts have been employed, the latter being based on noble metal (Pd, Pt, Rh) or earth-abundant metal complexes (Co, Ni, Fe, Mo) [4,5]. Among the latter, the catalytic properties of rhodium complexes have been rarely explored in aqueous solutions (Rh1–10) [3,6,7,8,9,10,11,12,13,14] but more largely in hydro-organic solvent (Rh11–24) [10,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33] (Scheme 1), although the use of water is a prerequisite for water-splitting applications [34]. Rhodium catalysts active in water for H2 production have employed a small variety of ligands, such as bipyridyl (Rh1–3, Rh8–9) [3,8,9,11,12,13], cyclopentadienyl (Rh3) [9], diphenylphosphinobenzene-sulphonate (Rh4) [6,7], acetate (Rh5–9) [8,10], and diisocyanopropane (Rh10) [8], under the form of mono- or bi-metallic complexes (Scheme 1a). The H2-evolution with such rhodium catalysts generally proceeds via the transient formation of rhodium hydride species [34,35,36] generated upon their reduction in presence of protons. Among these Rh catalysts, only Rh2, Rh3, and Rh4 exhibited turnover numbers per catalyst (TONCat) superior to 100 in the presence of a ruthenium or iridium photosensitizer and ascorbate as SD in water, and TONCat for the others catalysts do not exceed 10 [34]. In addition, when the Rh2 bis-bipyridyl catalyst is covalently linked to two Ru photosensitizers, the performance of photocatalytic system SD/PS-Cat is clearly improved due to the faster kinetics of the electron transfer between PS and Cat [12]. In this context, exploring rhodium complexes with other polypyridyl ligands is important for identifying new efficient and robust H2-evolving catalysts.
Herein, we report on the synthesis and structural characterization of a new rhodium mono-terpyridyl complex, namely [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (tpy = 2,2′;6′,2″-terpyridine) (denoted Rh-tpy, Scheme 2). We also investigated the electrochemical properties of this complex in organic solvent and the electrochemical generation and UV-visible spectroscopic characterization of the corresponding two-electron RhI reduced species. The catalytic activity of Rh-tpy towards the light-driven H2 production was tested in purely aqueous solution, in association with [Ru(bpy)3]2+ (Ru) as photosensitizer and ascorbate (HA) as sacrificial electron donor. In 2012, the group of Ogo investigated the capacity of the chemically synthetized rhodium(I) terpyridyl [RhI(tpy)(CH3CN)](CF3SO3) complex to form a rhodium(III) monohydride species in the presence of protons in CH3CN, [RhIII(H)(tpy)(CH3CN)2](CF3SO3)2, the latter being unstable and gradually transforming into a stable dinuclear RhII complex [Rh2II(tpy)2(CH3CN)4](CF3SO3)4, with the generation of H2 by reductive elimination [37]. In this study, we demonstrate that the Rh-tpy can act as a H2-evolving catalyst under photocatalytic conditions in acidic water (pH 4.0–4.5) with an efficiency competing with the benchmark rhodium(III) bis-dimethyl-bipyridyl complex [RhIII(dmbpy)2Cl2]+ (Rh2) (Scheme 1) for a low catalyst concentration of 1 μM, this latter being among the most efficient rhodium catalysts in water [11,12,13]. Finally, the first photo-induced steps involved in the H2 production with the molecular Ru/Rh-tpy/HA/H2A system are investigated by means of time-resolved emission and absorption spectroscopies.

2. Results and Discussion

2.1. Synthesis and Crystal Structure of the [RhIII(tpy)(CH3CN)Cl2](CF3SO3) Complex

The synthesis of the Rh-tpy complex was inspired from that of the [RhIII(tpy)(OH)(H2O)2](NO3)2 complex isolated by Ogo (Scheme 3) [38]. First, the [RhIII(tpy)Cl3] complex is generated by reacting RhCl3•3H2O with one equivalent of terpyridine in ethanol at 90 °C [39]. Then, in order to solubilize the rhodium complex in water for the light-driven H2 production experiments, a chloride ligand in axial position was exchanged by an acetonitrile molecule by reacting the tris-chloro complex with AgCF3SO3 in acetonitrile at 90 °C, leading to the formation of the [RhIII(tpy)(CH3CN)Cl2](CF3SO3) complex (Rh-tpy) with a yield of 50%. The water solubility of Rh-tpy is also favored by the presence of the triflate counter-anion.
Single crystals of Rh-tpy suitable for X-ray crystallography were obtained by diffusion of diethyl ether into an acetonitrile solution of the rhodium complex (Figure 1, Tables S1–S4). Rh-tpy crystallizes in the P1 space group and displays a distorted octahedral geometry around the rhodium atom. The three nitrogen atoms of the terpyridine and one chloride ligand are in the equatorial plane, while the nitrogen atom of CH3CN and one chloride ligand are in an axial position. The terpyridine ligand is quasi-planar, with low torsion angles between the central pyridine of the terpyridine and the other two pyridines (3.90° between pyridines with N(1) and N(2) and 1.70° between pyridines with N(2) and N(3)). The nitrogen atoms of the terpyridine, the chloride atom, and the rhodium form a slightly distorted square plane, with N(1)-Rh-N(2) and N(2)-Rh-N(3) angles of 81.07(6)° and 81.00(6)°, respectively, which are slightly smaller than the ideal value of 90° for a perfect square plane geometry. This distortion is induced by the terpyridine, which imposes 5-atom rings comprising the rhodium, two nitrogen atoms (N(1)/N(2) or N(2)/N(3)), and two terpyridine carbon atoms. The distance between rhodium and the central pyridine nitrogen (Ncentral) of the terpyridine (N(2)) is shorter (1.9386(14) Å) than that between rhodium and the two distal nitrogen (Ndistal) of the terpyridine (2.0258(14) and 2.0417(14) Å for Rh-N(1) and Rh-N(3), respectively). The rhodium atom is almost equidistant from the two chloride atoms (2.3530(5) vs. 2.3067(4) Å, respectively, for Rh-Cl(1) and Rh-Cl(2)) and the distance between the nitrogen of acetonitrile ligand and the rhodium atom (2.0220(14) Å for Rh-N(21)) is very similar to those between the rhodium and the terpyridine nitrogen atoms (Rh-Ntpy). The distances and angles between the Rh center and the terpyridine ligand in Rh-tpy are very similar to those found in the [RhIII(tpy)(OH)(OH2)2](NO3)2 [38]. Slightly shorter valence angles were found in the dinuclear [Rh2II(tpy)2(CH3CN)4](CF3SO3)4 complex where the two Rh centers are linked by a metal-metal bond, as a consequence of RhIII-Ntpy lengths being slightly longer than those of Rh-tpy. Regarding the low-valent [RhI(tpy)(CH3CN)](CF3SO3) complex, which adopts a discrete structure of four [RhI(tpy)(CH3CN)]+ units stacked via tpy-tpy interaction at the solid state, this complex displays slightly shorter valence angles and RhIII-Ntpy bond distances than those found in Rh-tpy [37].

2.2. Spectroelectrochemical Properties of the [RhIII(tpy)(CH3CN)Cl2](CF3SO3) Complex (Rh-tpy) in N,N-Dimethylformamide (DMF)

The redox properties of Rh-tpy were investigated in DMF because of the wider electrochemical window of this solvent compared to water, facilitating the observation of the reduced state of the rhodium complex (Figure 2 and Table S5). By analogy with the previously reported electrochemical behavior of the rhodium bis-bipyridyl [11,12,13,36,40] and bis-terpyridyl [41] complexes in organic solvent, the irreversible redox process at Epa = −1.13 V vs. Ag/AgNO3 observed on the cyclic voltammogram of Rh-tpy is attributed to the two-electron reduction of the metal center (RhIII  RhI), while the poorly reversible process located at a lower potential (Epc = −1.95 V) is ascribed to the one-electron reduction of the terpyridine ligand (Figure 2A). The irreversibility of the RhIII/RhI system is the result of the release of the two axial ligands upon reduction to form a RhI species with a square-planar geometry. On the reverse scan, the irreversible oxidation peak observed at Epa = −0.70 V, followed by another less defined at ca. −0.3 V, is related to the reoxidation of the metal center (RhI  RhIII) along with the recoordination of two exogeneous ligands, which can be the solvent (DMF), CH3CN, and/or a chloride ion.
In order to further characterize and evaluate the stability of the reduced RhI state of Rh-tpy, the initial species involved in the catalytic reduction of protons to H2, an electrolysis was performed at −1.25 V. In accordance with the generation of a RhI species, the exhaustive electrolysis required the exchange of two electrons per initial RhIII complex. The formation of this species was attested by the UV-Visible absorption spectrum (Figure 3) and cyclic voltammograms (Figure 2B) of the resulting electrolyzed solution. The initial pale yellow Rh-tpy solution displays mainly intense absorption bands below 350 nm. After complete reduction, the deep blue solution obtained from RhI exhibits two strong absorption bands at 279 and 327 nm, and four less intense absorption bands in the visible region at 384, 515, 607, and 665 nm (Figure 3). Very similar absorption spectra were reported by the group of Hartl in ethanol for the square-planar rhodium(I) terpyridyl halide complexes, [RhI(tpy)(X)] (X = Cl or Br), chemically synthesized from [Rh(X)(COD)]2 precursors (X = Cl or Br; COD = cycloocta-1,5-diene) under inert atmosphere [42]. The similarity of the absorption spectra strongly suggests the [RhI(tpy)Cl] structure for the electrogenerated RhI species as a result of the release of one CH3CN and one chloride ligand.
Regarding the cyclic voltammograms of the RhI solution, the initial irreversible reduction peak (RhIII → RhI) has fully disappeared and only the redox process centered on the tpy ligand is now present on the cyclic voltammogram, with an intensity similar to that of the initial solution of [RhIII(tpy)(MeCN)Cl2]+ (Figure 2A,B), attesting to the stability of the RhI species under inert atmosphere. In addition, the irreversible oxidation peaks (RhI → RhIII) observed by scanning from −1.1 to 0 V are accompanied, on the reverse scan, by the appearance of a RhIII → RhI reduction peak in accordance with the regeneration of a RhIII species. Besides, the UV-vis. spectrum and the cyclic voltammograms of the electrolyzed solution do not show any changes after several hours when the latter is stored under inert atmosphere free of oxygen (glove box), showing the excellent stability of the RhI terpyridyl complex. By contrast, under similar experimental conditions, solutions of [RhI(Rbpy)2]+ electrogenerated from rhodium(III) bis-bipyridyl complexes [RhIII(Rbpy)2Cl2]+ (R = dimethyl or ditertio-butyl) are less stable and start to decompose after 1.5 h [36]. This is likely due to the higher π-acceptor capacity of terpyridine compared to that of bipyridine, which discharges the excess of electron density of RhI to the ligand, thus stabilizing the low oxidation states of the metal [42]. The rigid and planar geometry of the tpy ligand can also contribute to the stability of the square planar RhI species. A back exhaustive electrolysis at E = 0 V restores a RhIII complex. However, the two-electron reduction of the metal center (RhIII → RhI) is now located at a slightly more negative potential (Epa = −1.27 V vs. Ag/AgNO3) compared to the initial solution of Rh-tpy, suggesting a slightly different coordination sphere for the RhIII center, which might be the substitution of the initial CH3CN axial ligand by DMF. Indeed, the two chloride ligands remain coordinated to the RhIII center, as judged by the absence of the characteristic electroactivity of free chlorides (irreversible oxidation peak at ca. 0.7 V vs. Ag/AgNO3) in the positive potential region from Rh-tpy [36].
This electrochemical study shows that the activation of the Rh-tpy complex for photo-induced H2 production is possible using Ru as photosensitizer. Indeed, Rh-tpy can be reduced either by the reduced [RuII(bpy)(bpy•−)]+ (denoted Ru) or by the excited states of Ru (denoted *RuII), since both reactions are favored from a thermodynamic point of view. Considering the reduction potential of Rh-tpy (Epc (RhIII/RhI) = −1.13 V vs. Ag/AgNO3 (−0.83 V vs. SCE)) in DMF, and the redox potential of Ru at reduced Ru and excited *RuII states in water (E1/2 (RuII/Ru) = −1.50 V and E1/2 (RuIII/*RuII) = −1.07 V vs. SCE [3], see Table S5), the driving forces (ΔG0) are clearly exergonic with values of −0.67 and −0.24 eV, respectively. Photophysical experiments (see below) allowed for deciphering the photocatalytic mechanism occurring with the system Ru/Rh-tpy/HA/H2A.

2.3. Photocatalytic Hydrogen Production

The catalytic activity of Rh-tpy for the light-driven production of H2 was evaluated in the presence of the Ru photosensitizer and ascorbate as a sacrificial electron donor in deaerated aqueous solutions (5 mL) at 298 K under visible light irradiation (400–700 nm). The light-driven evolution of H2 was determined by gas chromatography in the course of the photocatalysis, and these data were used to calculate the TON (turnover number) and initial TOF values (turnover frequency) per molecule of catalyst (respectively denoted as TONCat and TOFCat), which characterize the performance of the photocatalytic system (Table S7).
The catalytic activity of Rh-tpy was first investigated in a HA/H2A buffer (total concentration 1.1 M, see Table S6) at pH 4.0 with 530 μM of PS, similar to our previous experiments with Rh2 [11,12] and cobalt catalysts [43,44,45,46,47]. In these conditions, the Ru/Rh-tpy/HA/H2A system displays 220 TONCat after 22 h of irradiation with a catalyst concentration of 10 μM (Figure 4). However, the stability of the photocatalytic system did not exceed 7–8 h and about 90% of the total amount of H2 was achieved after 6 h. The UV-vis. absorption spectrum of the photocatalytic solution recorded at the end of the photocatalysis indicates the complete transformation of the Ru PS into the bis-bipyridyl Ru(II) derivative, [Ru(bpy)2(X)2]n+, with the disappearance of the initial absorption band at 450 nm of Ru and the appearance of a new shoulder at 473 nm with a lower intensity (Figure 5A). This is due to the well-known poor stability of the reduced state Ru in acidic water, generated upon the reductive quenching of *Ru by HA (see below), which easily undergoes a ligand substitution by solvent molecules and/or anions such as ascorbate [48,49]. In view of slowing down the degradation of Ru and thus improving the stability of the photocatalytic system, a lower HA/H2A concentration of 0.1 M was used. To keep the pH constant during the photocatalytic experiment, the aqueous solution was buffered with an acetate buffer (1 M) at pH 4.0. Such conditions have been already employed for several photocatalytic systems for H2 production involving cobalt catalysts [48,50,51,52,53]. In acetate buffer containing 0.1 M of HA/H2A at pH 4.0, the Ru/Rh-tpy photocatalytic system is clearly more stable, since the H2 production is still effective after 22 h of irradiation. At this stage, the TONCat value has reached 380, which is almost twice as high as the TONCat value obtained in 1.1 M HA/H2A buffer, and the Ru PS is less degraded at the end of photocatalysis (Figure 5B). However, when the concentration of the sacrificial electron donor is divided by 10, the initial TOFCat value is lower than that obtained in 1.1 M HA/H2A (38 vs. 95 TOFCat, respectively, see Table S7). The pH was further optimized using the acetate buffer with 0.1 M HA/H2A medium. The highest TONCat value of 423 was obtained at pH 4.5 after 22 h of irradiation. A further increase of the pH to 4.8 leads to a decrease of the TONCat value (Figure 4). Basically, the pH at which photocatalysis is optimal is governed by a delicate balance between the reactivity of the catalyst (generation of the RhIII hydride species) and the efficiency of the photoinduced electron transfer process (concentration of both protonated and non-protonated form of the SD) [34,54]. More acidic conditions make the protonation of the catalyst, i.e., the H2-evolution catalysis, faster, but the generation of the reductant Ru is slowed down because the concentration of the ascorbate electron donor is smaller. In other words, in less acidic conditions, the higher concentration of ascorbate makes the quenching or regeneration of PS more efficient, but the low concentration in protons decreases the H2-evolving activity of the catalyst.
In order to rule out the formation of Rh0 colloids from Rh-tpy during the photocatalysis, an experiment was carried out with 530 μM of Ru, 10 μM of Rh-tpy in acetate buffer (0.1 M) containing HA/H2A (0.1 M) at pH 4.5, in the presence of an excess of mercury (Figure 6). Rh0 colloids are known to catalyze the reduction of protons to H2 and can be deactivated by generating an amalgam with mercury [55,56]. The absence of effect on the photocatalytic behavior of the Ru/Rh-tpy/HA/H2A system due to the Hg addition confirms that no colloidal rhodium is formed during photocatalysis and the molecular Rh-tpy catalyst is involved in the H2 evolution.
The photocatalytic performances of the Ru/Rh-tpy/HA/H2A system were also investigated at lower concentrations of Rh-tpy in acetate buffer at pH 4.5 by keeping the concentration of Ru constant in view to promote the reduction of the catalyst (Figure 7). At 5 and 1 μM, the TONCat values increase drastically to reach 804 and 2242, respectively, after 22–25 h of irradiation. This strong increase of TONCat values when the PS/Cat ratio increases has been previously observed in many photocatalytic molecular systems using rhodium [11,15] or cobalt catalysts [45,46,47,48,49,53,57,58,59,60,61,62,63]. However, these values do not take into account the amount of H2 that comes only from the Ru photosensitizer in the absence of a catalyst. Indeed, while control experiments in the absence of Ru or HA do not produce any H2, some H2 is generated from solutions containing only Ru and HA/H2A (Table S7, Figure 7B). This amount is not negligible, as it represents about 30% of the H2 produced for the 10 and 5 μM concentrations and more than half (55%) for the lowest concentration of 1 μM. Consequently, the actual TONCat values (denoted as TONCat*), calculated by subtracting the amount of H2 stemming from Ru, are thus of 300, 558, and 1012 at 10, 5, and 1 μM, respectively (Table S7).
UV-vis absorption spectra recorded at the end of the photocatalysis show that the Ru PS is partially transformed in Ru-bis-bipyridine species and that this transformation is more pronounced as the concentration of the catalyst is low (Figure 8) [53]. The degradation of the PS is most probably the main limiting factor for the H2 production after 22 h [54].
The photocatalytic performance of Ru/Rh-tpy/HA/H2A has also been compared with that obtained with Rh2 as catalyst in the same experimental conditions (Table S7). In combination with a molecular photosensitizer, the Rh2 complex is among the most efficient H2 evolving catalysts based on rhodium that operate in water [11,12,15,34]. At pH 4.5 in 1 M acetate buffer containing 0.1 M of HA/H2A, the Rh2 catalyst appears significantly more active than Rh-tpy for the higher catalyst concentration of 10 μM, with corrected TONCat* of 1140 vs. 300 for Rh-tpy. However, in contrast to what was observed for Rh-tpy and, in a previous reported study for Rh2 in the 1.1 M HA/H2A buffer [11], if the catalyst concentration is decreased, the TONCat* also decreases to about 780 at 5 μM and 770 at 1 μM (Figure 7). Thus, Rh-tpy is clearly less efficient than Rh2 at the higher catalyst concentrations of 10 and 5 μM (TONCat* of 300 and 558 for Rh-tpy vs. 1140 and 778 for Rh2), but challenges Rh2 in more diluted conditions (for 1 μM, TONCat* of 1012 for Rh-tpy vs. 772 for Rh2). In fact, the higher stability of the RhI species for Rh-tpy compared to that of Rh2, as revealed in organic solvent, could explain the photocatalytic performance of Ru/Rh-tpy/HA/H2A competing with that of Ru/Rh2/HA/H2A at very low concentrations of catalyst. Therefore, the use of tridentate tpy ligand compared to bidentate bpy ligands leads to a less active but potentially more stable H2-evolving catalyst. However, the instability of Ru PS does not allow for comparing the long-term stability of Rh-tpy and Rh2 catalysts. A more stable PS in water such as the triazatriangulenium TATA+ organic dye should be employed, but the reduction potential of this PS, which is less negative than that of Ru [53], will not allow for an efficient reduction of the Rh catalysts.
We also summarized the performances of the various three-components photocatalytic systems for H2 production in homogeneous solution in Tables S8 and S9, involving a rhodium complex (Rh1–24, Scheme 1) as a H2-evolving catalyst previously reported in view to compare the efficiency of these systems with those of Rh2 and Rh-tpy as catalysts. Tables S8 and S9 summarize the performances in hydro-organic (Rh11–24, Scheme 1b) and in purely aqueous (Rh1–10, Scheme 1a) solution, respectively. The structures of the various photosensitizers employed in these photocatalytic systems are gathered in Scheme S1. In aqueous organic media, PSs mainly rely on heteroleptic cyclometaled iridium(III) complexes of the type [Ir(C^N)2(N^N)]+ (C^N = phenylpyridine ligand and N^N = bipyridine ligand) with various substituents, although there are a few examples employing copper. In purely aqueous solution, most of the systems rely on Ru. Reliable comparison of the efficiency of the different systems is not obvious, since in addition to the fact that the experimental conditions are carried out in different experimental set-up, TONCat values are strongly dependent on the PS and Cat concentrations, the PS/Cat ratio, the nature of the solvent and of sacrificial electron donor, and the long-term stability of the PS.
Among the Rh catalysts employed in water (Rh1–10, Scheme 1a), only Rh2, Rh3, and Rh4 exhibited turnover numbers per catalyst (TONCat) higher than 100 in the presence of a ruthenium or iridium photosensitizer and ascorbate as SD [11]; TONCat for the others catalysts, Rh5–10, do not exceed 10 [3,8] (Table S9). However, these later systems used significantly higher catalyst concentration ranging from 50 μM to 1.56 mM, and the catalyst concentration is generally higher than that of the PS, resulting in lower TONCat. Concerning the Wilkinson catalyst Rh4, although up to 2000 TONsCat were reported [6,7], further studies by our group have shown that this catalyst is much less efficient than Rh2 [11].
Significantly higher TONCat were obtained in hydro-organic media. The highest TONCat in such media, from 2362 up to 9886, were obtained by the group of Kataoka with the Rh5, Rh15–16 and Rh11 catalysts in association with Ir1, Ir2, and Ir7 as PSs in THF/H2O [28,29] (Table S8). In these examples, the concentration ranges are similar to our studies, i.e., 500 μM for the PS and 5 μM for the catalyst. Interestingly, in the other studies combining Ir3–5 (100 μM) and Rh1 or Rh11–14 (100 μM) of Bernhard [15], Ir1 (50 μM) and Rh8–9 or Rh17 (50 μM) of Wang [10] and Ir2 (926 μM) and Rh18–21 (926 μM) of Kozlova [32,33], for which the Ir/Cat ratio is close to 1/1, TONsCat are comprised between 15 and 150, and reach 380 in one case.
These examples show that even with quite similar families of catalysts (for instance, Rh5–9 [8] and Rh5, Rh15–16 [28,29]) and PS, the TONCat values can be very different depending on the experimental conditions. The most relevant comparisons are therefore those made under similar experimental conditions. We can thus compare the efficiency of our systems with rhodium with those of the cobalt(III) tetraazamacrocyclic [CoIII(CR14)Cl2]+ (Co) complex, which is one of the most efficient H2-evolving catalyst in acidic water [43,45,47,53]. The catalytic performances of this complex have been recorded under similar experimental conditions with 500 μM Ru and 0.1 M of HA/H2A in 1 M acetate buffer at pH 4.5 for catalyst concentrations of 10 and 5 μM (Table S7) [53]. TONCat* of 1086 and 1822 were obtained for Co compared to 300 and 558 for Rh-tpy for 10 and 5 μM, respectively. At both catalysts’ concentrations, Rh-tpy is thus much less efficient than Co, with more than three times less hydrogen produced.

2.4. Mechanistic Insight for the Ru/Rh-tpy/HA/H2A System from Photophysical Measurements

The light-driven H2 production with the Ru/Rh-tpy/HA/H2A three-component system is initiated by the generation of the excited state of the Ru, *RuII, under light absorption (Scheme 4). *Ru could be then quenched by an electron transfer to the sacrificial electron donor (HA) to generate the reduced form of PS, Ru, and the oxidized form of ascorbate (HA) (reductive quenching). Ru is able, in turn, to reduce the catalyst (RhIII-tpy to RhI-tpy), reforming the ground state Ru. HA can lose a proton to form A•−, which can disproportionate generating the dehydroascorbic acid (DHA) [34,64,65], a very good electron acceptor able to withdraw electrons from Ru and/or RhI-tpy. Therefore, a back electron transfer takes place from Ru to HA or DHA (BET process) to restore RuII and HA, or from RhI-tpy to HA or DHA (BETC process) to regenerate RhIII-tpy.
The *Ru can also be quenched by the catalyst leading to the oxidized form of the PS, RuIII, and the reduced state of catalyst, RhI-tpy (oxidative quenching), this process being unfavorable (see below).
Photophysical measurements allow us to identify which one of these two mechanisms is the most favorable. By using a Stern–Volmer plot, a rate constant (kQ1) of 1.0 × 107 M−1 s−1 was determined by the Schmehl’s group for the reductive quenching of the *RuII luminescence by HA in an aqueous acetate buffered at pH 4.5. In the same medium, we determined a rate constant of 6.87 × 108 M−1 s−1 for the oxidative quenching (kQ2) of *Ru by Rh-tpy (Figure 9), which is about 70 times higher than that of the reductive quenching of *Ru by HA. Although the oxidative quenching is kinetically more favored than the reductive quenching, the reductive pathway dominates over the oxidative one with pseudo-first-order kinetics of 1.0 × 106 s−1 and 0.069–3.4 × 104 s−1, respectively, considering that the HA concentration (0.1 M) is much higher than that of Rh-tpy (1–10 μM) under photocatalytic conditions. Noteworthily, the rate constant of the oxidative quenching of *Ru by RhIII-tpy is very similar to that previously determined by our group between *Ru and Rh2 (3.2 × 108 M−1 s−1) [11]. This similarity of rate constants could be correlated to their akin driving forces ΔG0 of −0.24 eV for *Ru/Rh-tpy and −0.28 eV for *Ru/Rh2, considering the reduction potentials (RhIII/RhI) of Rh-tpy and Rh2, and the oxidation potential of *Ru (see Table S5).
Nanosecond flash photolysis experiments have also been performed to characterize the photoinduced electron transfer process occurring in the system Ru/Rh-tpy/HA/H2A at pH 4.5. In the absence of Rh-tpy, the transient absorption spectra recorded after excitation at 455 nm (Figure S1) show the formation of the Ru species with positive absorption bands at 360 and 510 nm [66]. The formation of Ru occurs from an electron transfer between *Ru and HA [9,11]. Ru growth follows a pseudo first-order kinetics with an estimated rate constant of 4.3 × 106 s−1 (τ = 235 ns, Figure S2). The decay of the transient absorption trace at 510 nm, due to the back electron transfer (BET) between Ru and the oxidized forms of ascorbate (HA and DHA), can be fitted according to a second order kinetics law (Figure S3). The rate constant (kBET) is estimated to be 7.4 × 109 M−1 s−1.
In the presence of Rh-tpy, flash photolysis experiments also show the formation of the Ru species, evidenced by an increase of absorption at 360 and 510 nm (Figure 10). The growth of the signal occurs in a similar time scale (τ = 215 ns, corresponding to a constant at 4.6 × 106 s−1, Figure S4) to that observed without Rh-tpy (Figure S2). This confirms that the photocatalytic cycle is initiated by a reductive quenching of *Ru by HA and that the presence of Rh-tpy does not interfere with this first electron transfer process. However, the decay of the transient absorption trace at 510 nm is accelerated in the presence of Rh-tpy (Figure 11). This is a consequence of an electron transfer process between the transient Ru and RhIII-tpy leading to the initial RuII and RhI-tpy, which efficiently competes with the BET process between Ru and the oxidized ascorbate (e.g., HA and DHA).
In the presence of Rh-tpy, the decay of the transient Ru species is best fitted by a pseudo-first-order kinetics law, leading to a rate constant for the ET process of 2.4 × 104 s−1 (Figure S5). Considering the concentration of the catalyst in solution (200 μM), this would correspond to a bimolecular rate constant (kET) of 1.2 × 108 M−1 s−1, which is 62 times lower than the kBET value (7.4 × 109 M−1 s−1). The difference between these two kinetics should contradict the faster decay of Ru in presence of RhIII-tpy observed in the transient absorption trace in Figure 11. Nevertheless, we have estimated the initial concentration of Ru (2.27 × 10−7 M), calculated from the concentration of *RuII (9.1 × 10−7 M) just after the excitation pulse (ΔA@450 nm = −0.01 and Δε = −11,000 M−1 cm−1) and considering a quantum yield of 25% for the Ru formation from reaction between *Ru and HA [45]. This has allowed us to determine the real rates of the ET and BET processes (vET and vBET) at 5.45 × 10−3 and 3.81 × 10−4 M s−1, respectively (see ESI for the calculation details). In other words, under our experimental conditions, the ET process dominates over the BET process with a vET/vBET ratio of 14.3. Finally, although the two-electron reduced catalyst (RhI-tpy) exhibits a large absorption band between 550 and 700 nm (Figure 3), its typical spectroscopic signature is not observed in the transient absorption spectra. This is attributed to the fast reactivity of the RhI-tpy species (i.e., [RhI(tpy)X]n+, X = Cl or H2O with n = 0 or 1, respectively) with protons, below the nanosecond time-scale, to form a RhIII(H)-tpy hydride species (i.e., [RhIII(H)(tpy)(H2O)X]n+, X = Cl or H2O with n = 1 or 2, respectively), the key intermediate to reduce protons into H2. Indeed, from this species, different pathways can lead to the production of H2 via homolytic (reaction with another RhIII(H)-tpy species to generate H2 and two Rh(II) species) or heterolytic (reaction with a proton to form H2 and a Rh(III) species) route (Scheme 4). RhIII(H)-tpy could be also further reduced by Ru to a Rh(II) hydride species, RhII(H)-tpy, from which H2 can be released via similar homolytic and heterolytic pathways, generating Rh(I) and Rh(II) species, respectively. The group of Ogo has shown that in CH3CN, the hydride complex [RhIII(H)(tpy)(CH3CN)2]2+ can slowly generate H2 via reductive elimination leading to the Rh(II) dimer species (homolytic route) [37]. However, we cannot rule out that in acidic water, other pathways can proceed in parallel to this homolytic mechanism. For instance, according to theoretical calculations, we have shown that, for the Rh2 catalyst in acidic water, H2 is preferentially released through a heterolytic mechanism from the RhIII(H) species and that both homo- and heterolytic mechanisms are thermodynamically favorable to generate H2 via the RhII(H) species [13]. Furthermore, although the formation of the RhIII(H)-tpy hydride species is very fast, the catalysis of proton reduction generally remains the rate-determining step of the photocatalytic mechanism, occurring within three-component systems for photocatalytic evolution of H2 in homogeneous media [44,54].

3. Materials and Methods

3.1. Synthesis of [RhIII(tpy)Cl3]

RhCl3•3H2O (150 mg, 0.55 mmol) and 2, 2′:6′; 2″-terpyridine (129 mg, 0.55 mmol) were dissolved in absolute ethanol (24 mL). After stirring at 90 °C for 12 h, a yellow precipitate appeared to correspond to [RhIII(tpy)Cl3], which was filtered off from the hot reaction mixture and washed with diethyl ether (188 mg, yield, 77%). 1H NMR (DMSO-d6), 400 MHz, δ (ppm): 9.28 (d; J = 5.2 Hz; 2H); 8.81 (d; J = 8 Hz; 2H); 8.77 (d; J = 7.6 Hz; 2H); 8.54 (t; J = 8 Hz; 1H); 8.38 (t; J = 7.6 Hz; 2H); 7.95 (t; J = 6.4 Hz; 2H). IR (cm−1): 3069, 2932, 1598, 1570, 1472, 1447, 1396, 1312, 1252, 1241, 1184, 1157, 1136, 1048, 1026, 751, 668, 655. ESI-MS, positive mode m/z (%): 405.87 (59) [M − Cl]+, 463.83 (81) [M + Na]+.

3.2. Synthesis of [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy)

[RhIII(tpy)Cl3] (116 mg, 0.26 mmol) was added to an acetonitrile solution (232 mL) of AgCF3SO3 (209 mg, 0.81 mmol). After stirring at 90 °C for 20 h, the white precipitate of AgCl formed was filtered off from the hot reaction mixture. The resulting pale yellow filtrate was concentrated until 30 mL, yielding to a yellow precipitate of [RhIII(tpy)(CH3CN)Cl2](CF3SO3), which was filtered off and washed with diethyl ether to give a pale yellow powder. [RhIII(tpy)(CH3CN)Cl2](CF3SO3) was recrystallized by slow diffusion of diethyl ether into an acetonitrile solution of complex (80 mg, yield, 51%). 1H NMR (D2O), 400 MHz, δ (ppm): 9.06 (d; J = 5.6 Hz; 2H); 8.66 (d; J = 7.6 Hz; 2H); 8.59 (d; J = 7 Hz; 1H); 8.46 (t; J = 8 Hz; 2H); 8.02 (t; J = 7.6 Hz; 2H); 2.94 (s, 3H); IR: 3088, 2930, 2324, 1604, 1575, 1479, 1452, 1403, 1320, 1258, 1223, 1139, 1030, 773, 635. ESI-MS, positive mode m/z (%): 446.95 (100) [M − CF3SO3]+, 405.93 (26) [M − CF3SO3-CH3CN]+.

4. Conclusions

We demonstrated, for the first time, that the terpyridyl rhodium complex, [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy), is able to catalyze the light-induced proton reduction into H2 in aqueous solution in the presence of the [RuII(bpy)3]Cl2 (Ru) photosensitizer and ascorbate as sacrificial electron donors. Under photocatalytic conditions, TONCat values of 300 and up to 1000 were obtained for H2 production for a Rh-tpy catalyst concentration at 10 and 1 μM, respectively, after subtraction of the amount of H2 stemming from the Ru only. The photocatalytic performance of Ru/Rh-tpy/HA/H2A has also been compared with that obtained with Rh2 as a catalyst in the same experimental conditions. It appears that Rh-tpy is clearly less efficient than Rh2 at the higher catalyst concentrations of 10 and 5 μM (TONCat* of 300 and 558 for Rh-tpy vs 1140 and 778 for Rh2), but challenges Rh2 in more diluted conditions (for 1 μM, TONCat* of 1012 for Rh-tpy vs. 772 for Rh2). Therefore, the use of tridentate tpy ligand compared to bidentate bpy ligands leads to a less active but potentially more stable H2-evolving catalyst. However, Rh-tpy, is much less efficient than the cobalt(III) tetraazamacrocyclic [CoIII(CR14)Cl2]+ (Co) complex, ione of the most efficient H2-evolving catalysts in acidic water, since with this catalyst, TONCat* of 1086 and 1822 were reached at 10 and 5 μM, respectively.
The electrochemical study in DMF reveals that the reduced state of the rhodium catalyst, identified as [RhI(tpy)Cl] (RhI-tpy) from its UV-visible signature, is stable for several hours under an inert atmosphere owing to the π-acceptor property of the terpyridine ligand that stabilizes the low oxidation states of the rhodium. A good stability of the low-valent form of the rhodium catalyst makes it less prone to degradation in the course of photocatalysis. The π-acceptor property of terpyridine also confers moderately negative reduction potential to the Rh-tpy catalyst (about −0.8 V vs. SCE), making its reduction by the reduced state of Ru effective (E1/2(RuII/RuI) = −1.50 V vs. SCE). A Stern–Volmer plot and transient absorption spectroscopy have shown that the first step of the photocatalytic process is a reductive quenching of the Ru excited state by ascorbate. The resulting Ru species is then able to reduce the RhIII-tpy catalyst generating RhI-tpy, which reacts with a proton on a sub-nanosecond time scale to form a RhIII(H)-tpy hydride, the key intermediate for H2 evolution. The search for new molecular catalysts for H2 production with simple and easily synthesized ligands is still ongoing, and the terpyridine ligand, with its particular electronic and coordination properties, is a good candidate for designing new catalysts to meet these requirements.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196614/s1, Materials and general experimental details. Additional data and experimental details for X-ray structure determination, electrochemistry, UV-Vis absorption spectroscopy, 1H NMR, mass spectrum, photocatalytic hydrogen production, photophysics, nanosecond transient absorption spectroscopy (Figures S1–S8 and Tables S1–S7). CCDC 2173038 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 [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; fax: +44-1223-336033.

Author Contributions

Conceptualization, M.-N.C. and J.F.; methodology, M.-N.C., J.F., F.C. and D.A.; validation, F.C., M.-N.C. and J.F.; formal analysis, F.C. and J.F.; investigation, F.C., T.G., B.D., J.C. and J.P.; resources, M.-N.C. and J.F.; data curation, F.C. and J.F.; writing—original draft preparation, M.-N.C. and J.F.; writing—review and editing, M.-N.C.; visualization, M.-N.C.; supervision, M.-N.C. and J.F.; project administration, M.-N.C. and J.F.; funding acquisition, M.-N.C. and J.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been partially supported by the French National Research Agency through the project TATADyes (ANR-20-CE05-0041) for the Master1 grant of T.G., Labex ARCANE and CBH-EUR-GS (ANR-17-EURE-0003) and by “Université Grenoble Alpes” for the PhD grants of F.C. and B.D.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to the CNRS, the Université Grenoble Alpes and the French National Research Agency through the project TATADyes (ANR-20-CE05-0041), Labex ARCANE and CBH-EUR-GS (ANR-17-EURE-0003) for their financial supports. This work was also supported by ICMG Chemistry Nanobio Platform (FR2067). F.C. and B.D. thank the “Université Grenoble Alpes” for their PhD grant, and T.G. thanks the French National Research Agency (TATADyes; ANR-20-CE05-0041) for his Master1 grant.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds [RhIII(tpy)Cl3] and [RhIII(tpy)(CH3CN)Cl2](CF3SO3) are available from the authors.

References

  1. Armaroli, N.; Balzani, V. The Hydrogen Issue. ChemSusChem 2011, 4, 21–36. [Google Scholar] [CrossRef] [PubMed]
  2. Lehn, J.M.; Sauvage, J.P. Chemical Storage of Light energy—Catalytic Generation of Hydrogen by Visible-Light or Sunlight Irradiation of Neutral Aqueous Solution. Nouv. J. Chim. 1977, 1, 449–451. [Google Scholar]
  3. Kirch, M.; Lehn, J.M.; Sauvage, J.P. Hydrogen Generation by Visible-Light Irradiation of Aqueous-Solutions of Metal-Complexes—Approach to the Photo-Chemical Conversion and Storage of Solar-Energy. Helv. Chim. Acta 1979, 62, 1345–1384. [Google Scholar] [CrossRef]
  4. Eckenhoff, W.T.; Eisenberg, R. Molecular systems for light driven hydrogen production. Dalton Trans. 2012, 41, 13004–13021. [Google Scholar] [CrossRef]
  5. Dalle, K.E.; Warnan, J.; Leung, J.J.; Reuillard, B.; Karmel, I.S.; Reisner, E. Electro- and Solar-Driven Fuel Synthesis with First Row Transition Metal Complexes. Chem. Rev. 2019, 119, 2752–2875. [Google Scholar] [CrossRef]
  6. Oishi, S. A Water-Soluble Wilkinsons Complex as Homogeneous Catalyst for the Photochemical Reduction of Water. J. Mol. Catal. 1987, 39, 225–232. [Google Scholar] [CrossRef]
  7. Bauer, R.; Werner, H.A.F. Investigations on a Homogeneous Wilkinsons Catalyst for the Water Photolysis. Int. J. Hydrogen Energy 1994, 19, 497–499. [Google Scholar] [CrossRef]
  8. Tanaka, S.; Masaoka, S.; Yamauchi, K.; Annaka, M.; Sakai, K. Photochemical and thermal hydrogen production from water catalyzed by carboxylate-bridged dirhodium(II) complexes. Dalton Trans. 2010, 39, 11218–11226. [Google Scholar] [CrossRef]
  9. Fukuzumi, S.; Kobayashi, T.; Suenobu, T. Photocatalytic Production of Hydrogen by Disproportionation of One-Electron-Reduced Rhodium and Iridium–Ruthenium Complexes in Water. Angew. Chem. Int. Ed. 2011, 50, 728–731. [Google Scholar] [CrossRef]
  10. Xie, J.; Li, C.; Zhou, Q.; Wang, W.; Hou, Y.; Zhang, B.; Wang, X. Large Improvement in the Catalytic Activity Due to Small Changes in the Diimine Ligands: New Mechanistic Insight into the Dirhodium(II,II) Complex-Based Photocatalytic H2 Production. Inorg. Chem. 2012, 51, 6376–6384. [Google Scholar] [CrossRef]
  11. Stoll, T.; Gennari, M.; Serrano, I.; Fortage, J.; Chauvin, J.; Odobel, F.; Rebarz, M.; Poizat, O.; Sliwa, M.; Deronzier, A.; et al. [RhIII(dmbpy)2Cl2]+ as a Highly Efficient Catalyst for Visible-Light-Driven Hydrogen Production in Pure Water: Comparison with Other Rhodium Catalysts. Chem.-Eur. J. 2013, 19, 782–792. [Google Scholar] [CrossRef]
  12. Stoll, T.; Gennari, M.; Fortage, J.; Castillo, C.E.; Rebarz, M.; Sliwa, M.; Poizat, O.; Odobel, F.; Deronzier, A.; Collomb, M.-N. An Efficient RuII–RhIII–RuII Polypyridyl Photocatalyst for Visible-Light-Driven Hydrogen Production in Aqueous Solution. Angew. Chem. Int. Ed. 2014, 53, 1654–1658. [Google Scholar] [CrossRef]
  13. Kayanuma, M.; Stoll, T.; Daniel, C.; Odobel, F.; Fortage, J.; Deronzier, A.; Collomb, M.-N. A computational mechanistic investigation of hydrogen production in water using the [RhIII(dmbpy)2Cl2]+/[RuII(bpy)3]2+/ascorbic acid photocatalytic system. Phys. Chem. Chem. Phys. 2015, 17, 10497–10509. [Google Scholar] [CrossRef]
  14. Canterbury, T.R.; Arachchige, S.M.; Brewer, K.J.; Moore, R.B. A new hydrophilic supramolecular photocatalyst for the production of H2 in aerobic aqueous solutions. Chem. Commun. 2016, 52, 8663–8666. [Google Scholar] [CrossRef]
  15. 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. [Google Scholar] [CrossRef]
  16. Miyake, Y.; Nakajima, K.; Sasaki, K.; Saito, R.; Nakanishi, H.; Nishibayashi, Y. Design and Synthesis of Diphosphine Ligands Bearing an Osmium(II) Bis(terpyridyl) Moiety as a Light-Harvesting Unit: Application to Photocatalytic Production of Dihydrogen. Organometallics 2009, 28, 5240–5243. [Google Scholar] [CrossRef]
  17. Wang, J.; White, T.A.; Arachchige, S.M.; Brewer, K.J. A new structural motif for photoinitiated electron collection: Ru,Rh bimetallics providing insight into H2 production via photocatalysis of water reduction by Ru,Rh,Ru supramolecules. Chem. Commun. 2011, 47, 4451–4453. [Google Scholar] [CrossRef]
  18. White, T.A.; Higgins, S.L.H.; Arachchige, S.M.; Brewer, K.J. Efficient Photocatalytic Hydrogen Production in a Single-Component System Using Ru,Rh,Ru Supramolecules Containing 4,7-Diphenyl-1,10-Phenanthroline. Angew. Chem. Int. Ed. 2011, 50, 12209–12213. [Google Scholar] [CrossRef]
  19. White, T.A.; Whitaker, B.N.; Brewer, K.J. Discovering the Balance of Steric and Electronic Factors Needed To Provide a New Structural Motif for Photocatalytic Hydrogen Production from Water. J. Am. Chem. Soc. 2011, 133, 15332–15334. [Google Scholar] [CrossRef]
  20. Peuntinger, K.; Pilz, T.D.; Staehle, R.; Schaub, M.; Kaufhold, S.; Petermann, L.; Wunderlin, M.; Görls, H.; Heinemann, F.W.; Li, J.; et al. Carbene based photochemical molecular assemblies for solar driven hydrogen generation. Dalton Trans. 2014, 43, 13683–13695. [Google Scholar] [CrossRef] [Green Version]
  21. Zhou, R.; Manbeck, G.F.; Wimer, D.G.; Brewer, K.J. A new RuIIRhIII bimetallic with a single Rh–Cl bond as a supramolecular photocatalyst for proton reduction. Chem. Commun. 2015, 51, 12966–12969. [Google Scholar] [CrossRef] [PubMed]
  22. Wagner, A.T.; Zhou, R.; Quinn, K.S.; White, T.A.; Wang, J.; Brewer, K.J. Tuning the Photophysical Properties of Ru(II) Monometallic and Ru(II),Rh(III) Bimetallic Supramolecular Complexes by Selective Ligand Deuteration. J. Phys. Chem. A 2015, 119, 6781–6790. [Google Scholar] [CrossRef] [PubMed]
  23. Mengele, A.K.; Kaufhold, S.; Streb, C.; Rau, S. Generation of a stable supramolecular hydrogen evolving photocatalyst by alteration of the catalytic center. Dalton Trans. 2016, 45, 6612–6618. [Google Scholar] [CrossRef] [PubMed]
  24. Kagalwala, H.N.; Chirdon, D.N.; Mills, I.N.; Budwal, N.; Bernhard, S. Light-Driven Hydrogen Generation from Microemulsions Using Metallosurfactant Catalysts and Oxalic Acid. Inorg. Chem. 2017, 56, 10162–10171. [Google Scholar] [CrossRef] [PubMed]
  25. Manbeck, G.F.; Fujita, E.; Brewer, K.J. Tetra- and Heptametallic Ru(II),Rh(III) Supramolecular Hydrogen Production Photocatalysts. J. Am. Chem. Soc. 2017, 139, 7843–7854. [Google Scholar] [CrossRef] [PubMed]
  26. McCullough, B.J.; Neyhouse, B.J.; Schrage, B.R.; Reed, D.T.; Osinski, A.J.; Ziegler, C.J.; White, T.A. Visible-Light-Driven Photosystems Using Heteroleptic Cu(I) Photosensitizers and Rh(III) Catalysts To Produce H2. Inorg. Chem. 2018, 57, 2865–2875. [Google Scholar] [CrossRef] [PubMed]
  27. Saeedi, S.; Xue, C.; McCullough, B.J.; Roe, S.E.; Neyhouse, B.J.; White, T.A. Probing the Diphosphine Ligand’s Impact within Heteroleptic, Visible-Light-Absorbing Cu(I) Photosensitizers for Solar Fuels Production. ACS Appl. Energy Mater. 2019, 2, 131–143. [Google Scholar] [CrossRef]
  28. Kataoka, Y.; Yano, N.; Handa, M.; Kawamoto, T. Intrinsic hydrogen evolution capability and a theoretically supported reaction mechanism of a paddlewheel-type dirhodium complex. Dalton Trans. 2019, 48, 7302–7312. [Google Scholar] [CrossRef]
  29. Kataoka, Y.; Yano, N.; Kohara, Y.; Tsuji, T.; Inoue, S.; Kawamoto, T. Experimental and Theoretical Study of Photochemical Hydrogen Evolution Catalyzed by Paddlewheel-Type Dirhodium Complexes with Electron Withdrawing Carboxylate Ligands. ChemCatChem 2019, 11, 6218–6226. [Google Scholar] [CrossRef]
  30. Saeedi, S.; White, T.A. Insight into the Reductive Quenching of a Heteroleptic Cu(I) Photosensitizer for Photocatalytic H2 Production. ACS Appl. Energy Mater. 2020, 3, 56–65. [Google Scholar] [CrossRef] [Green Version]
  31. Whittemore, T.J.; Xue, C.; Huang, J.; Gallucci, J.C.; Turro, C. Single-chromophore single-molecule photocatalyst for the production of dihydrogen using low-energy light. Nat. Chem. 2020, 12, 180–185. [Google Scholar] [CrossRef]
  32. Vasilchenko, D.; Tkachev, S.; Baidina, I.; Korolkov, I.; Berdyugin, S.; Kozlova, E.; Kozlov, D. Preparation of a rhodium(iii) cis-diaquacomplex by protic acid induced oxalate-release from mer-[Rh(C2O4)Cl(py)3]. New J. Chem. 2018, 42, 19637–19643. [Google Scholar] [CrossRef]
  33. Vasilchenko, D.B.; Tkachev, S.V.; Kurenkova, A.Y.; Kozlova, E.A.; Kozlov, D.V. Photocatalytic hydrogen evolution by iridium(III)–rhodium(III) system: Effect of the free ligand addition. Int. J. Hydrogen Energy 2016, 41, 2592–2597. [Google Scholar] [CrossRef]
  34. Stoll, T.; Castillo, C.E.; Kayanuma, M.; Sandroni, M.; Daniel, C.; Odobel, F.; Fortage, J.; Collomb, M.-N. Photo-induced redox catalysis for proton reduction to hydrogen with homogeneous molecular systems using rhodium-based catalysts. Coord. Chem. Rev. 2015, 304–305, 20–37. [Google Scholar] [CrossRef]
  35. Bakac, A. Aqueous rhodium(III) hydrides and mononuclear rhodium(II) complexes. Dalton Trans. 2006, 13, 1589–1596. [Google Scholar] [CrossRef]
  36. Castillo, C.E.; Stoll, T.; Sandroni, M.; Gueret, R.; Fortage, J.; Kayanuma, M.; Daniel, C.; Odobel, F.; Deronzier, A.; Collomb, M.-N. Electrochemical Generation and Spectroscopic Characterization of the Key Rhodium(III) Hydride Intermediates of Rhodium Poly(bipyridyl) H2-Evolving Catalysts. Inorg. Chem. 2018, 57, 11225–11239. [Google Scholar] [CrossRef]
  37. Inoki, D.; Matsumoto, T.; Nakai, H.; Ogo, S. Experimental Study of Reductive Elimination of H2 from Rhodium Hydride Species. Organometallics 2012, 31, 2996–3001. [Google Scholar] [CrossRef]
  38. Inoki, D.; Matsumoto, T.; Hayashi, H.; Takashita, K.; Nakai, H.; Ogo, S. Establishing the mechanism of Rh-catalysed activation of O2 by H2. Dalton Trans. 2012, 41, 4328–4334. [Google Scholar] [CrossRef]
  39. Paul, P.; Tyagi, B.; Bilakhiya, A.K.; Bhadbhade, M.M.; Suresh, E.; Ramachandraiah, G. Synthesis and Characterization of Rhodium Complexes Containing 2,4,6-Tris(2-pyridyl)-1,3,5-triazine and Its Metal-Promoted Hydrolytic Products: Potential Uses of the New Complexes in Electrocatalytic Reduction of Carbon Dioxide. Inorg. Chem. 1998, 37, 5733–5742. [Google Scholar] [CrossRef]
  40. Castillo, C.E.; Gennari, M.; Stoll, T.; Fortage, J.; Deronzier, A.; Collomb, M.N.; Sandroni, M.; Légalité, F.; Blart, E.; Pellegrin, Y.; et al. Visible Light-Driven Electron Transfer from a Dye-Sensitized p-Type NiO Photocathode to a Molecular Catalyst in Solution: Toward NiO-Based Photoelectrochemical Devices for Solar Hydrogen Production. J. Phys. Chem. C 2015, 119, 5806–5818. [Google Scholar] [CrossRef]
  41. Paul, J.; Spey, S.; Adams, H.; Thomas, J.A. Synthesis and structure of rhodium complexes containing extended terpyridyl ligands. Inorg. Chim. Acta 2004, 357, 2827–2832. [Google Scholar] [CrossRef]
  42. de Pater, B.C.; Frühauf, H.-W.; Vrieze, K.; de Gelder, R.; Evert, J.B.; McCormack, D.; Lutz, M.; Anthony, L.S.; Hartl, F. Strongly Nucleophilic RhI Centre in Square-Planar Complexes with Terdentate (κ3) 2,2′:6′,2′′-Terpyridine Ligands: Crystallographic, Electrochemical and Density Functional Theoretical Studies. Eur. J. Inorg. Chem. 2004, 2004, 1675–1686. [Google Scholar] [CrossRef] [Green Version]
  43. Varma, S.; Castillo, C.E.; Stoll, T.; Fortage, J.; Blackman, A.G.; Molton, F.; Deronzier, A.; Collomb, M.-N. Efficient photocatalytic hydrogen production in water using a cobalt(III) tetraaza-macrocyclic catalyst: Electrochemical generation of the low-valent Co(I) species and its reactivity toward proton reduction. Phys. Chem. Chem. Phys. 2013, 15, 17544–17552. [Google Scholar] [CrossRef]
  44. Gimbert-Surinach, C.; Albero, J.; Stoll, T.; Fortage, J.; Collomb, M.-N.; Deronzier, A.; Palomares, E.; Llobet, A. Efficient and Limiting Reactions in Aqueous Light-Induced Hydrogen Evolution Systems using Molecular Catalysts and Quantum Dots. J. Am. Chem. Soc. 2014, 136, 7655–7661. [Google Scholar] [CrossRef] [PubMed]
  45. Gueret, R.; Castillo, C.E.; Rebarz, M.; Thomas, F.; Hargrove, A.-A.; Pécaut, J.; Sliwa, M.; Fortage, J.; Collomb, M.-N. Cobalt(III) tetraaza-macrocyclic complexes as efficient catalyst for photoinduced hydrogen production in water: Theoretical investigation of the electronic structure of the reduced species and mechanistic insight. J. Photochem. Photobiol. B Biol. 2015, 152, 82–94. [Google Scholar] [CrossRef] [PubMed]
  46. Lo, W.K.C.; Castillo, C.E.; Gueret, R.; Fortage, J.; Rebarz, M.; Sliwa, M.; Thomas, F.; McAdam, C.J.; Jameson, G.B.; McMorran, D.A.; et al. Synthesis, Characterization, and Photocatalytic H2-Evolving Activity of a Family of [Co(N4Py)(X)]n+ Complexes in Aqueous Solution. Inorg. Chem. 2016, 55, 4564–4581. [Google Scholar] [CrossRef] [PubMed]
  47. Gueret, R.; Castillo, C.E.; Rebarz, M.; Thomas, F.; Sliwa, M.; Chauvin, J.; Dautreppe, B.; Pécaut, J.; Fortage, J.; Collomb, M.-N. Cobalt(II) Pentaaza-Macrocyclic Schiff Base Complex as Catalyst for Light-Driven Hydrogen Evolution in Water: Electrochemical Generation and Theoretical Investigation of the One-Electron Reduced Species. Inorg. Chem. 2019, 58, 9043–9056. [Google Scholar] [CrossRef] [PubMed]
  48. Singh, W.M.; Baine, T.; Kudo, S.; Tian, S.L.; Ma, X.A.N.; Zhou, H.Y.; DeYonker, N.J.; Pham, T.C.; Bollinger, J.C.; Baker, D.L.; et al. Electrocatalytic and Photocatalytic Hydrogen Production in Aqueous Solution by a Molecular Cobalt Complex. Angew. Chem. Int. Ed. 2012, 51, 5941–5944. [Google Scholar] [CrossRef]
  49. 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. [Google Scholar] [CrossRef] [Green Version]
  50. Singh, W.M.; Mirmohades, M.; Jane, R.T.; White, T.A.; Hammarstrom, L.; Thapper, A.; Lomoth, R.; Ott, S. Voltammetric and Spectroscopic Characterization of Early Intermediates in the Co(II)-Polypyridyl-Catalyzed Reduction of Water. Chem. Commun. 2013, 49, 8638–8640. [Google Scholar] [CrossRef]
  51. Natali, M.; Badetti, E.; Deponti, E.; Gamberoni, M.; Scaramuzzo, F.A.; Sartorel, A.; Zonta, C. Photoinduced hydrogen evolution with new tetradentate cobalt(ii) complexes based on the TPMA ligand. Dalton Trans. 2016, 45, 14764–14773. [Google Scholar] [CrossRef]
  52. Lucarini, F.; Pastore, M.; Vasylevskyi, S.; Varisco, M.; Solari, E.; Crochet, A.; Fromm, K.M.; Zobi, F.; Ruggi, A. Heptacoordinate CoII Complex as a New Architecture for Photochemical Hydrogen Production. Chem. Eur. J. 2017, 23, 6768–6771. [Google Scholar] [CrossRef] [Green Version]
  53. Gueret, R.; Poulard, L.; Oshinowo, M.; Chauvin, J.; Dahmane, M.; Dupeyre, G.; Lainé, P.P.; Fortage, J.; Collomb, M.-N. Challenging the [Ru(bpy)3]2+ Photosensitizer with a Triazatriangulenium Robust Organic Dye for Visible-Light-Driven Hydrogen Production in Water. ACS Catal. 2018, 8, 3792–3802. [Google Scholar] [CrossRef]
  54. Costentin, C.; Camara, F.; Fortage, J.; Collomb, M.-N. Photoinduced Catalysis of Redox Reactions. Turnover Numbers, Turnover Frequency, and Limiting Processes: Kinetic Analysis and Application to Light-Driven Hydrogen Production. ACS Catal. 2022, 12, 6246–6254. [Google Scholar] [CrossRef]
  55. Amouyal, E.; Koffi, P. Photochemical Production of Hydrogen from Water. J. Photochem. 1985, 29, 227–242. [Google Scholar] [CrossRef]
  56. Weddle, K.S.; Aiken, J.D.; Finke, R.G. Rh(0) nanoclusters in benzene hydrogenation catalysis: Kinetic and mechanistic evidence that a putative [(C8H17)3NCH3]+ [RhCl4] ion-Pair catalyst is actually a distribution of Cl and [(C8H17)3NCH3]+ stabilized Rh(0) nanoclusters. J. Am. Chem. Soc. 1998, 120, 5653–5666. [Google Scholar] [CrossRef]
  57. Sandroni, M.; Gueret, R.; Wegner, K.D.; Reiss, P.; Fortage, J.; Aldakov, D.; Collomb, M.-N. Cadmium-Free CuInS2/ZnS Quantum Dots as Efficient and Robust Photosensitizers in combination with a Molecular Catalyst for Visible Light-Driven H2 Production in Water. Energy Environ. Sci. 2018, 11, 1752–1761. [Google Scholar] [CrossRef]
  58. 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. [Google Scholar] [CrossRef]
  59. Guttentag, M.; Rodenberg, A.; Bachmann, C.; Senn, A.; Hamm, P.; Alberto, R. A highly stable polypyridyl-based cobalt catalyst for homo- and heterogeneous photocatalytic water reduction. Dalton Trans. 2013, 42, 334–337. [Google Scholar] [CrossRef]
  60. Deponti, E.; Luisa, A.; Natali, M.; Iengo, E.; Scandola, F. Photoinduced hydrogen evolution by a pentapyridine cobalt complex: Elucidating some mechanistic aspects. Dalton Trans. 2014, 43, 16345–16353. [Google Scholar] [CrossRef]
  61. Natali, M.; Luisa, A.; Lengo, E.; Scandola, F. Efficient photocatalytic hydrogen generation from water by a cationic cobalt(II) porphyrin. Chem. Commun. 2014, 50, 1842–1844. [Google Scholar] [CrossRef]
  62. 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. [Google Scholar] [CrossRef]
  63. Vennampalli, M.; Liang, G.; Katta, L.; Webster, C.E.; Zhao, X. Electronic Effects on a Mononuclear Co Complex with a Pentadentate Ligand for Catalytic H2 Evolution. Inorg. Chem. 2014, 53, 10094–10100. [Google Scholar] [CrossRef]
  64. Bachmann, C.; Probst, B.; Guttentag, M.; Alberto, R. Ascorbate as an electron relay between an irreversible electron donor and Ru(II) or Re(I) photosensitizers. Chem. Commun. 2014, 50, 6737–6739. [Google Scholar] [CrossRef] [PubMed]
  65. Guttentag, M.; Rodenberg, A.; Kopelent, R.; Probst, B.; Buchwalder, C.; Brandstätter, M.; Hamm, P.; Alberto, R. Photocatalytic H2 Production with a Rhenium/Cobalt System in Water under Acidic Conditions. Eur. J. Inorg. Chem. 2012, 2012, 59–64. [Google Scholar] [CrossRef]
  66. Mulazzani, Q.G.; Emmi, S.; Fuochi, P.G.; Hoffman, M.Z.; Venturi, M. On the nature of tris(2,2′-bipyridine)ruthenium(1+) ion in aqueous solution. J. Am. Chem. Soc. 1978, 100, 981–983. [Google Scholar] [CrossRef]
Scheme 1. Rhodium catalysts previously reported for light-driven H2 production in water (a) and in hydro-organic solvent (b).
Scheme 1. Rhodium catalysts previously reported for light-driven H2 production in water (a) and in hydro-organic solvent (b).
Molecules 27 06614 sch001
Scheme 2. Structures of the photosensitizer (Ru), the H2-evolving catalyst (Rh-tpy), and the sacrificial electron donor (HA) used in this work.
Scheme 2. Structures of the photosensitizer (Ru), the H2-evolving catalyst (Rh-tpy), and the sacrificial electron donor (HA) used in this work.
Molecules 27 06614 sch002
Scheme 3. Synthesis of the [RhIII(tpy)Cl3] (a) and [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (b) (Rh-tpy) Complexes.
Scheme 3. Synthesis of the [RhIII(tpy)Cl3] (a) and [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (b) (Rh-tpy) Complexes.
Molecules 27 06614 sch003
Figure 1. ORTEP diagram of [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy). Thermal ellipsoids are drawn at the 50% probability level. Selected distances (Å) and angles (º): Rh1-N1 = 2.0258(14) Å, Rh1-N2 = 1.9386(14) Å, Rh1-N3 = 2.0417(14) Å, Rh1-N21 = 2.0220(14) Å, Rh1-Cl1 = 2.3530(5) Å, Rh1-Cl2 = 2.3067(4) Å, N1-Rh1-N3 = 162.03(6)º, N2-Rh1-N1 = 81.07(6)º, N2-Rh1-N3 = 81.00(6)º, N2-Rh1-N21 = 87.90(6)º.
Figure 1. ORTEP diagram of [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy). Thermal ellipsoids are drawn at the 50% probability level. Selected distances (Å) and angles (º): Rh1-N1 = 2.0258(14) Å, Rh1-N2 = 1.9386(14) Å, Rh1-N3 = 2.0417(14) Å, Rh1-N21 = 2.0220(14) Å, Rh1-Cl1 = 2.3530(5) Å, Rh1-Cl2 = 2.3067(4) Å, N1-Rh1-N3 = 162.03(6)º, N2-Rh1-N1 = 81.07(6)º, N2-Rh1-N3 = 81.00(6)º, N2-Rh1-N21 = 87.90(6)º.
Molecules 27 06614 g001
Figure 2. Cyclic voltammograms at a glassy carbon electrode (Ø = 3 mm, ν = 100 mV s−1) of (A) 0.5 mM solution of Rh-tpy in DMF, 0.1 M [Bu4N]ClO4, (B) after a two-electron exhaustive reduction at –1.25 V vs. Ag/AgNO3 leading to the formation of the Rh(I) species, and (C) after a back exhaustive oxidation at 0 V.
Figure 2. Cyclic voltammograms at a glassy carbon electrode (Ø = 3 mm, ν = 100 mV s−1) of (A) 0.5 mM solution of Rh-tpy in DMF, 0.1 M [Bu4N]ClO4, (B) after a two-electron exhaustive reduction at –1.25 V vs. Ag/AgNO3 leading to the formation of the Rh(I) species, and (C) after a back exhaustive oxidation at 0 V.
Molecules 27 06614 g002
Figure 3. UV-vis absorption spectra of 0.5 mM solution of Rh-tpy in DMF, 0.1 M [Bu4N]ClO4 before (green trace) and after a two-electron exhaustive reduction at −1.25 V vs. Ag/AgNO3 leading to the formation of the [RhI(tpy)Cl] (blue trace), followed by a back exhaustive oxidation at 0 V (orange trace). Optical path length was 1 mm.
Figure 3. UV-vis absorption spectra of 0.5 mM solution of Rh-tpy in DMF, 0.1 M [Bu4N]ClO4 before (green trace) and after a two-electron exhaustive reduction at −1.25 V vs. Ag/AgNO3 leading to the formation of the [RhI(tpy)Cl] (blue trace), followed by a back exhaustive oxidation at 0 V (orange trace). Optical path length was 1 mm.
Molecules 27 06614 g003
Figure 4. Photocatalytic hydrogen production (TONCat) as a function of time under 400–700 nm irradiation from a deaerated aqueous solution (5 mL) of Ru (530 μM) and Rh-tpy (10 μM) containing an acetate buffer (1 M) and NaHA/H2A (0.1 M) at different pHs (4.0; 4.5; 4.8) (solid trace) or only a buffer of NaHA/H2A (1.1 M) at pH 4.0 (dotted trace).
Figure 4. Photocatalytic hydrogen production (TONCat) as a function of time under 400–700 nm irradiation from a deaerated aqueous solution (5 mL) of Ru (530 μM) and Rh-tpy (10 μM) containing an acetate buffer (1 M) and NaHA/H2A (0.1 M) at different pHs (4.0; 4.5; 4.8) (solid trace) or only a buffer of NaHA/H2A (1.1 M) at pH 4.0 (dotted trace).
Molecules 27 06614 g004
Figure 5. UV-vis absorption spectra of (A) Ru (530 μM) in aqueous solution containing NaHA/H2A (1.1 M) at pH 4.0 in presence of Rh-tpy (10 μM), (B) Ru (530 μM) in acetate buffer (1 M) at pH 4.0 containing NaHA/H2A (0.1 M) in presence of Rh-tpy (10 μM), initial solutions and after about 22 h of irradiation. Optical path = 1 mm.
Figure 5. UV-vis absorption spectra of (A) Ru (530 μM) in aqueous solution containing NaHA/H2A (1.1 M) at pH 4.0 in presence of Rh-tpy (10 μM), (B) Ru (530 μM) in acetate buffer (1 M) at pH 4.0 containing NaHA/H2A (0.1 M) in presence of Rh-tpy (10 μM), initial solutions and after about 22 h of irradiation. Optical path = 1 mm.
Molecules 27 06614 g005
Figure 6. Photocatalytic hydrogen production (TONCat) as a function of time under 400–700 nm irradiation from deaerated 1 M acetate aqueous buffer (5 mL) at pH 4.5 containing NaHA/H2A (0.1 M), in the presence of Ru (530 μM) and Rh-tpy (10 μM) with (black trace) and without (red trace) one drop of mercury.
Figure 6. Photocatalytic hydrogen production (TONCat) as a function of time under 400–700 nm irradiation from deaerated 1 M acetate aqueous buffer (5 mL) at pH 4.5 containing NaHA/H2A (0.1 M), in the presence of Ru (530 μM) and Rh-tpy (10 μM) with (black trace) and without (red trace) one drop of mercury.
Molecules 27 06614 g006
Figure 7. Photocatalytic hydrogen production as a function of time under 400–700 nm irradiation from a deaerated 1 M acetate aqueous buffer (5 mL) at pH 4.5 containing NaHA/H2A (0.1 M), in the presence of Ru (530 μM) and Rh-tpy at 10, 5, and 1 μM: (A) TONCat and (B) nH2 and VH2.
Figure 7. Photocatalytic hydrogen production as a function of time under 400–700 nm irradiation from a deaerated 1 M acetate aqueous buffer (5 mL) at pH 4.5 containing NaHA/H2A (0.1 M), in the presence of Ru (530 μM) and Rh-tpy at 10, 5, and 1 μM: (A) TONCat and (B) nH2 and VH2.
Molecules 27 06614 g007
Figure 8. UV-vis absorption spectra of Ru (530 μM) in acetate buffer (1 M) at pH 4.5 and NaHA/H2A (0.1 M), initial solution and after about 22 h of irradiation with Rh-tpy at various concentrations (10, 5, and 1 μM). Optical path = 1 mm.
Figure 8. UV-vis absorption spectra of Ru (530 μM) in acetate buffer (1 M) at pH 4.5 and NaHA/H2A (0.1 M), initial solution and after about 22 h of irradiation with Rh-tpy at various concentrations (10, 5, and 1 μM). Optical path = 1 mm.
Molecules 27 06614 g008
Scheme 4. Proposed catalytic mechanism for the light-driven H2 production with the system Ru/Rh-tpy/HA/H2A. For reasons of simplicity, only the pathway showing the heterolytic mechanism from RhIII(H)-tpy is shown. *RuII represents the triplet excited state of RuII, and Ru, its one-electron reduced state.
Scheme 4. Proposed catalytic mechanism for the light-driven H2 production with the system Ru/Rh-tpy/HA/H2A. For reasons of simplicity, only the pathway showing the heterolytic mechanism from RhIII(H)-tpy is shown. *RuII represents the triplet excited state of RuII, and Ru, its one-electron reduced state.
Molecules 27 06614 sch004
Figure 9. Stern–Volmer plot performed in a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 for the oxidative quenching of Ru (10 μM) at an excited state by Rh-tpy by varying its concentration (0; 0.06; 0.08; 0.1; 0.15; 0.2; 0.3 mM). The lifetime of *Ru without catalyst (τ0) was determined to be 620 ns in water at pH 4.5.
Figure 9. Stern–Volmer plot performed in a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 for the oxidative quenching of Ru (10 μM) at an excited state by Rh-tpy by varying its concentration (0; 0.06; 0.08; 0.1; 0.15; 0.2; 0.3 mM). The lifetime of *Ru without catalyst (τ0) was determined to be 620 ns in water at pH 4.5.
Molecules 27 06614 g009
Figure 10. Transient absorption spectra recorded at different times after flash laser excitation (λ = 455 nm) of a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 containing Ru (100 μM), NaHA/H2A (0.1 M), and Rh-tpy (200 μM) (path length = 1 cm).
Figure 10. Transient absorption spectra recorded at different times after flash laser excitation (λ = 455 nm) of a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 containing Ru (100 μM), NaHA/H2A (0.1 M), and Rh-tpy (200 μM) (path length = 1 cm).
Molecules 27 06614 g010
Figure 11. Transient absorption traces on the microsecond time scale recorded at 510 nm after laser excitation at 455 nm of a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 containing: (a) Ru (100 μM) and NaHA/H2A (0.1 M), (b) Ru (100 μM), NaHA/H2A (0.1 M), and Rh-tpy (200 μM).
Figure 11. Transient absorption traces on the microsecond time scale recorded at 510 nm after laser excitation at 455 nm of a deaerated aqueous solution of acetate buffer (1 M) at pH 4.5 containing: (a) Ru (100 μM) and NaHA/H2A (0.1 M), (b) Ru (100 μM), NaHA/H2A (0.1 M), and Rh-tpy (200 μM).
Molecules 27 06614 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Camara, F.; Gavaggio, T.; Dautreppe, B.; Chauvin, J.; Pécaut, J.; Aldakov, D.; Collomb, M.-N.; Fortage, J. Electrochemical Properties of a Rhodium(III) Mono-Terpyridyl Complex and Use as a Catalyst for Light-Driven Hydrogen Evolution in Water. Molecules 2022, 27, 6614. https://doi.org/10.3390/molecules27196614

AMA Style

Camara F, Gavaggio T, Dautreppe B, Chauvin J, Pécaut J, Aldakov D, Collomb M-N, Fortage J. Electrochemical Properties of a Rhodium(III) Mono-Terpyridyl Complex and Use as a Catalyst for Light-Driven Hydrogen Evolution in Water. Molecules. 2022; 27(19):6614. https://doi.org/10.3390/molecules27196614

Chicago/Turabian Style

Camara, Fakourou, Thomas Gavaggio, Baptiste Dautreppe, Jérôme Chauvin, Jacques Pécaut, Dmitry Aldakov, Marie-Noëlle Collomb, and Jérôme Fortage. 2022. "Electrochemical Properties of a Rhodium(III) Mono-Terpyridyl Complex and Use as a Catalyst for Light-Driven Hydrogen Evolution in Water" Molecules 27, no. 19: 6614. https://doi.org/10.3390/molecules27196614

Article Metrics

Back to TopTop