Next Article in Journal
Transformation of Dilute Ethylene at High Temperature on Micro- and Nano-Sized H-ZSM-5 Zeolites
Next Article in Special Issue
Electrodeposition of Fe-Complexes on Oxide Surfaces for Efficient OER Catalysis
Previous Article in Journal
A Focus on the Transformation Processes for the Valorization of Glycerol Derived from the Production Cycle of Biofuels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Na3[Ru2(µ-CO3)4] as a Homogeneous Catalyst for Water Oxidation; HCO3 as a Co-Catalyst

1
Chemical Sciences Department and The Radical Research Center, Ariel University, Ariel 40700, Israel
2
Department of Chemistry, Arunachal University of Studies, NH52, Namsai, Arunachal Pradesh 792103, India
3
Chemistry Department, Nuclear Research Centre Negev, Beer-Sheva 84105, Israel
4
Chemistry Department, Ben-Gurion University, Beer-Sheva 84190, Israel
*
Authors to whom correspondence should be addressed.
Submission received: 30 January 2021 / Revised: 14 February 2021 / Accepted: 18 February 2021 / Published: 21 February 2021

Abstract

:
In neutral medium (pH 7.0) [RuIIIRuII(µ-CO3)4(OH)]4− undergoes one electron oxidation to form [RuIIIRuIII(µ-CO3)4(OH)2]4− at an E1/2 of 0.85 V vs. NHE followed by electro-catalytic water oxidation at a potential ≥1.5 V. When the same electrochemical measurements are performed in bicarbonate medium (pH 8.3), the complex first undergoes one electron oxidation at an Epa of 0.86 V to form [RuIIIRuIII(µ-CO3)4(OH)2]4−. This complex further undergoes two step one electron oxidations to form RuIVRuIII and RuIVRuIV species at potentials (Epa) 1.18 and 1.35 V, respectively. The RuIVRuIII and RuIVRuIV species in bicarbonate solutions are [RuIVRuIII(µ-CO3)4(OH)(CO3)]4− and [RuIVRuIV(µ-CO3)4(O)(CO3)]4− based on density functional theory (DFT) calculations. The formation of HCO4 in the course of the oxidation has been demonstrated by DFT. The catalyst acts as homogeneous water oxidation catalyst, and after long term chronoamperometry, the absorption spectra does not change significantly. Each step has been found to follow a proton coupled electron transfer process (PCET) as obtained from the pH dependent studies. The catalytic current is found to follow linear relation with the concentration of the catalyst and bicarbonate. Thus, bicarbonate is involved in the catalytic process that is also evident from the generation of higher oxidation peaks in cyclic voltammetry. The detailed mechanism has been derived by DFT. A catalyst with no organic ligands has the advantage of long-time stability.

Graphical Abstract

1. Introduction

With the decrease in the abundance of fossil fuels (coal, petroleum, and natural gas), the search for alternative energy sources is an immense challenge for mankind [1]. In the past few decades, solar energy and electricity have been considered as the source of alternative energies [2]. It is important to note that in nature, plants employ the highly sophisticated machinery called Photosystem II to convert sunlight into fuel [3]. It uses the calcium-manganese based oxo-cluster (CaMn4O5 core) as the catalyst to split water via a series of proton-coupled electron transfer (PCET) processes [4]. Because of pollution free combustion and high energy density, hydrogen is considered a green and sustainable source of energy that can be produced by splitting water using solar photocatalysis [5] or electrocatalysis [6]. For this, an efficient and durable catalyst is in high demand for promoting proton coupled evolution of oxygen with the removal of four electrons (Equation (1)). However, this process is sluggish and needs electrochemically high overpotential to oxidize the water.
2H2O → O2 + 4H+ + 4eE0 = 1.23 V vs. NHE
This is due to the complexity of the process: Formation of the O-O bond along with four protons and four electrons, which results in slow kinetics and thermodynamics [7]. RuO2 and IrO2 serve as highly efficient heterogeneous water oxidation catalysts [8]. Apart from oxides, sulfide, nitride, and phosphide transition metal compounds have been shown as catalysts of the oxygen evolution reaction, OER [9].
Over the years, extensive research has been carried out for the development of heterogeneous [10] and homogeneous [11] OER catalysts based on earth-abundant low cost materials. In the literature, a plethora of materials based on Mn [12,13], Fe [14,15], Co [15,16,17,18,19,20], Ni [21,22], and Cu [23], can be found which act as promising water oxidation catalysts. Catalysts involving costly metal, e.g., Ru [24] and Ir [25], are also reported. In the past few decades, enormous efforts have been put forwarded for the ruthenium-based molecular water oxidation catalysts [26,27]. Meyer and coworker [28] first reported the polypyridyl ruthenium-based water oxidation catalyst; cis, cis-[{RuII(bpy)2(py)(H2O)}2O]4+ (where py = pyridine and bpy = bipyridine) widely known as the “blue dimer” (BD). BD has been considered as the landmark in the mechanistic study of water oxidation catalysts (WOC) [29]. It has been proposed and verified by various in situ experimental studies that before OER the catalyst goes through a dimeric oxo species, O=RuV–RuV=O [30,31,32,33,34]. Three pathways have been proposed for the formation of the O-O bond after the water nucleophilic attack (Equations (3)–(5)) via coupling interaction of two metallo–oxyl/hydroxyl radicals are proposed [31,35].
L1Mn꞊O + H2O → L1Mn − 2−OOH + H+
2L1Mn꞊O → L1Mn − 1−OO−Mn − 1L1
L1Mn−(OH)2 → L1Mn − 2−(O2) + 2H+
2L1Mn−OH → L1Mn − 1−OO−Mn − 1L1 + 2H+
Later on, this mechanism of O-O bond formation became widely accepted by means of many experimental studies on molecular water oxidation catalysts [27,36]. Another important aspect of the BD catalyst in OER is that it involves PCETs [10,37,38,39,40]. PCETs are important because of two reasons: (a) Nature employs it in Photosystem II [41] and (b) it can decrease the pKa of the H2O molecule bound to the central cation, as a result, the formation of hydroxo/oxo species becomes easier, which stabilizes higher oxidation states (IV, V) [29]. However, in this context, it should be noted that the true role of homogeneous WOCs based on transition metal ions supported by organic/inorganic ligands are in question because in many cases it has been found that those complexes act as precursors for the generation of more reactive nanoparticles on the electrode surface [7].
Carbonate/bicarbonate acts as co-catalysts of the water oxidation processes [42,43]. This is due to the oxidation potential of the CO3•−/CO32− (E0 = 1.57 V vs. NHE) [44,45] couple that is significantly lower compared to that of OH/H2O (E0 = 2.73 V vs. NHE) or OH/OH [42]. Bicarbonate/carbonate can facilitate the water oxidation by two ways: Stabilizing the metal ions in high oxidation states due to its strong σ donor character and/or act as pro-oxidant by forming CO3•− and/or HCO4/C2O62−. For example, the redox potential of FeIII/II [46], MnIII/II [47], and CeIV/III [48] in concentrated carbonate solutions is shifted cathodically by 1.09, 0.99, and 1.70 V, respectively, compared to that of their aqueous counterpart. Similarly, other lanthanide carbonates, MIV(CO3)n (M = Pr, Tb), can be easily prepared by electrolysis of the carbonate solution at 1.4 V [48]. In 1958, Warburg and Krippahl reported that CO2 acts as stimulator in the light driven water oxidation process in Photosystem II [49]. Later on, it was accepted as the “bicarbonate effect” by means of experimental studies on the role of CO2/HCO4 [50].
The first experimental proof of the OER by copper carbonate was reported by Meyer et al. [51]. It showed the dependence of the catalytic current on metal salt concentration as well as on [CO32−]/[HCO3], but did not confirm the oxidation state of the active catalyst (CuIII or CuIV) in the rate determining step. Later, a Density Functional Theory (DFT) study by Cramer et al. [52] showed the possibility of a CuIV species as the active intermediate. However, Meyerstein and coworkers [53] have shown that CuIII(CO3)n3−2n is the active intermediate in the electrolysis of CuII(CO3)n2−2n solutions using pulse-radiolysis experiments and DFT studies. In a recent paper, the role of carbonate as a “proton shuttle” has been discussed in the electrocatalytic water oxidation by Cu(N,N′-2,6-dimethylphenyl-2,6-pyridinedicarbox-amidate), CuL [54] in the presence of carbonate. In another nickel complex, NiII(1,4,8,11-tetraazacyclotetra- decane)2+, NiIIL12+ the involvement of carbonate has been discussed by Burg et al. [22] where the formation of a NiIV=O active species via cleavage of the C-O bond of the carbonate has been postulated along with many other processes involving CO32−/HCO3. The formation of peroxo-mono-carbonate in the electro-catalytic water oxidation by some aluminum porphyrin complexes has been reported very recently by Kuttassery et al. [55] in the presence of CO32−/HCO3. Several mechanisms for the role of carbonate in these processes can be proposed [22,45,54,56]:
Mn(OH)(OCO2)l → (OCO2)l-1Mn − 2(OOCO2) + H+
Mn(O)(OCO2)l → (OCO2)l-1Mn − 2(OOCO2)
Mn(OCO2)l → Mn − 2(OCO2)l-2 + C2O62− (C2O62− + H+ → HCO4 + CO2)
Mn(OCO2)l → Mn(O)(OCO2)l-1 + CO2
Followed by reactions (6), (9), (10) or (11) [22]:
2Mn(O)(OCO2)l-1 → (OCO2)l-1Mn − 1-OO-Mn − 1(OCO2)l-1
Mn(O)(OCO2)l-1 + H2O → Mn − 2(OOH)(OCO2)l-1 + H+
Mn(O)(OCO2)l-1 + HCO3 → Mn − 2(OCO2)l-1 + HCO4
Mn(OH)(OCO2)l-1 + HCO3 → Mn − 2(OCO2)l-1 + HCO4 + H+
Mn(OCO2)l + HCO3 → Mn − 2(OCO2)l-1 + C2O62− + H+
2Mn(OCO2)l → 2Mn − 1(OCO2)l-1 + C2O62−
Mn(OCO2)l + HCO3 → Mn − 1(OCO2)l-1 + C2O62− + H+ + e (at an anode)
Mn(OH)(OCO2)l + HCO3 → Mn − 1(OCO2)l-1 + HCO4 + H+ + e (at an anode)
Mn(OCO2)l + H2O → Mn − 1(OCO2)l-1 + HCO4 + H+ + e (at an anode)
Mn(OH)(OCO2)l + H2O → Mn − 1(OCO2)l + H2O2 + H+ + e (at an anode)
As discussed above, the ruthenium based WOCs are promising and have interesting mechanisms of OER activity based on mononuclear and binuclear catalysts. The carbonate complex of ruthenium, Na3[Ru2(μ-CO3)4], has been reported by Wilkinson and coworkers [57,58] in 1986, and some electro-chemical properties have been studied by Cotton and coworkers [59]. Since water molecules can be coordinated to the two ruthenium central cations, in higher oxidation states, and this complex has reasonable solubility, though, most of the transition metal carbonates have low solubility in water, and in aqueous solution it seemed reasonable to speculate that it would be an ideal catalyst for electrochemical water oxidation. Herein the OER activity of the complex in the presence and absence of bicarbonate/carbonate are reported. Interestingly, in the presence of bicarbonate/carbonate, the catalyst goes through various oxidation states as obtained from cyclic voltammetry.

2. Results

The dimeric complex of ruthenium, Na3[Ru2(µ-CO3)4], is obtained as an orange yellow precipitate by reacting [Ru2(µ-CH3COO)4Cl] with Na2CO3 [59]. X-ray crystal structure [59] shows that four carbonates are bridging the two Ru centers to form the dimer, and the free oxygens of the carbonate ligands are bound to the axial position of a Ru atom present in an another dimer, making a two dimensional coordination polymeric structure (Figure S1). In the solid-state, no water molecule is found to be bound to the axial position of the ruthenium atoms. Thus, it has been considered that in solution the complex exists as [Ru2(µ-CO3)4]3−. The Ru25+ core has metal-metal (d6−d5) bonding with an electronic configuration of σ2π4δ2δ*2π*1 and bond order of 2.5 [60]. One interesting aspect of this complex is that, with further oxidation of the complex, electrons are being removed from the antibonding orbitals, resulting in an increase in bond order and strength. As a result, the integrity of the dimeric structure remains intact and is not decomposed. Herein two aspects of catalytic water oxidation are discussed: (a) Catalytic water oxidation in neutral medium and (b) catalytic water oxidation in bicarbonate/carbonate media.

2.1. Catalytic Water Oxidation in Neutral Medium

The complex was first dissolved in water in the presence of 0.20 M NaClO4 at a pH of 7.0 and cyclic voltammetry was performed. The quasi reversible (ΔEp = EpaEpc = 107 mV) oxidation has been observed at an E1/2 of 0.85 V vs. NHE (Figure 1).
We speculated that in this medium, the high valent site, RuIII, of the complex gets coordinated by OH. This fact has also been supported by theoretical calculations that show that when the complex is coordinated by OH it gets stabilized by a free energy gain of −102.32 kcal.mol−1. The one electron oxidation of the complex from [RuIIRuIII(µ-CO3)4(OH)]4− to [RuIIIRuIII(µ-CO3)4(OH)2]4− is confirmed by comparing the current with K4[Fe(CN)6] under similar conditions (Figure S2). In this context, it is important to note that Cotton and coworkers [59] have studied the electrochemical properties of this system and found the same oxidation potential for the RuIIIRuIII/RuIIIRuII couple. However, herein we show that if we scan the potential further, then a large current can be seen at potentials ≥ 1.5 V vs. NHE. A very weak peak can also be seen at 1.4 V, which indicates the catalytic process goes through high oxidation states of the metal. The first peak is diffusion controlled, as it follows the Randles–Sevcik equation (Equation (20)), i.e., the current is proportional to the square root of the scan rate (Figure 2 and Figure S3).
i d = 0.496 n d   α 1 2 F A [ C ] ( n d F v D o R T ) 1 2
The diffusion coefficient, Do, is calculated to be 1.16 × 10−6 cm2·s−1 from the slope of id vs. v1/2. Using Nicholson’s method [61], one calculates the rate constant (k0) of electron transfer to be (1.15 ± 0.15) × 10−3 cm·s−1 (Table S1). Herein, before water oxidation, the catalyst undergoes oxidation to various other high oxidation states and finally the O=RuIV–RuIV=O species is formed (see below), which acts as the active intermediate in the OER. However, other oxidation states can only be seen in cyclic voltammetry with large catalytic currents in the presence of bicarbonate/carbonate.
It is assumed that in the first step in neutral medium [RuIIRuIII(µ-CO3)4(OH)]4− undergoes one electron oxidation and simultaneously coordinates one OH to the open axial position of the other RuIII ion to form [RuIIIRuIII(µ-CO3)4(OH)2]4−, (Equation (21)).
[RuIIRuIII(µ-CO3)4(OH)]4− + H2O → [RuIIIRuIII(µ-CO3)4(OH)2]4− + H+ + e (pH 7.0)
This is manifested by the pH dependence of the anodic peak potential (Epa). The Epa shifts cathodically with the increase in pH with a slope of −50 mV·pH−1 (Figure 3 and Figure S4). The increase in current is due to a contribution from the next step. Further, the water oxidation peak potential also shifts linearly with the pH of the medium (Figure 4 and Figure S5). Thus, protons are involved in the rate determining step of the water oxidation.

2.2. Catalytic Water Oxidation in Bicarbonate/Carbonate Medium

The complex was dissolved in 0.10 M NaHCO3 and then cyclic voltammetric measurements were performed at pH 8.3. The cyclic voltammograms of the complex in the presence and absence of NaHCO3 are compared in Figure 5. Unlike the neutral medium (where only one prominent redox process is observed before the catalytic process), in the presence of bicarbonate, three redox processes before the catalytic water oxidation can be clearly seen. To further study the processes before the catalytic process, a square wave voltammogram was recorded (Figure 6). Interestingly four electrochemical processes are observed: (a) 1st peak due to the redox couple RuIIIRuIII/RuIIIRuII involving one electron at an Epa of 0.86 V vs. NHE, (b) 2nd peak due to the redox couple RuIVRuIII/RuIIIRuIII involving one electron oxidation at 1.18 V, (c) 3rd peak due to the redox couple RuIVRuIV/RuIVRuIII involving one electron oxidation at 1.35 V (little bit of large current probably due to a contribution from the next step), and (d) large current due to oxidation of HCO3/water at potential ≥ 1.5 V vs. NHE. Next, we shall discuss each step and the mechanisms involved in the process.

2.2.1. The First Wave, RuIIIRuIII/RuIIIRuII Redox Couple

The first quasi-reversible oxidation is observed at 0.86 V and is attributed to the oxidation of [RuIIRuIII(µ-CO3)4(OH)]4− to [RuIIIRuIII(µ-CO3)4(OH)2]4−. In the alkaline medium, the current of the 1st oxidation peak seems to increase, this increase is due to the overlap with the next electrochemical process (Figure S6). The 1st step is a one electron process (RuIIRuIII → RuIIIRuIII). It is important to note that in neutral medium, [RuIIRuIII(µ-CO3)4(OH)]4− undergoes one electron oxidation to form [RuIIIRuIII(µ-CO3)4(OH)2]4− (Equation (21)). The redox potential of the couple, RuIIIRuIII/RuIIIRuII, decreases consistently with increasing pH (Figure 7 and Figure S7).
The slope of the plot of Epa of RuIIIRuIII/RuIIIRuII vs. pH is −71 mV·pH−1, which is consistent with Equation (21). It should be noted that the increase in the pH also increases the ratio [CO32−]/[HCO3]. The scan rate dependence of the first two redox peaks are shown in Figure S8. Diffusion coefficient (Do = 0.51 × 10−6 cm2 s−1) decreases in bicarbonate medium. Further, to check the effect of bicarbonate concentration on the first redox wave, the CVs of the first redox wave were recorded in various bicarbonate concentrations. However, no significant change in the peak potential/current was found (Figure S9). This is because bicarbonate replaces the hydroxide ion after the redox process and is not involved in the electron transfer step.

2.2.2. The Second Wave, RuIVRuIII/RuIIIRuIII Redox Couple

The second wave is observed at an E1/2 of 0.97 V vs. NHE and is attributed to the oxidation of [RuIIIRuIII(µ-CO3)4(OH)2]4− to [RuIVRuIII(µ-CO3)4(OH)(CO3)]4− (Equation (22)) with the simultaneous release of one proton, i.e., a PCET process is taking place. The proposal that when the Ru center is oxidized to the +4 oxidation state the axial OH is replaced by a carbonate ligand is based on the DFT calculations, see below. The involvement of one electron is based on the comparison of the peak current
[RuIIIRuIII(µ-CO3)4(OH)2]4− + HCO3 → RuIVRuIII(µ-CO3)4(OH)(CO3)]4− + H2O + e
with K4[Fe(CN)6] under identical conditions (Figure S6). The peak current for the second wave is 7.6 µA, which is comparable to the current (7.1 µA) of FeIII/II couple in ferrocyanide. The pH dependence (Figure S7) of the peak potential confirms the PCET process. However, the slope of the peak potential (Epa) of the [RuIVRuIII(µ-CO3)4(OH)(CO3)]4−/[RuIIIRuIII(µ-CO3)4(OH)2]4− is only −18 mV·pH−1 (Figure 8) against the theoretical value of −59 mV·pH−1. However, the partial overlap with the third electrochemical process (Figure S7), which increases with the pH, might cause this. This peak is quasi-reversible (ΔEp = 124 mV), and the reduction peak can be seen in the reverse scan.

2.2.3. The Third Wave, RuIVRuIV/RuIVRuIII Redox Couple

The third peak is observed at 1.35 V and is due to a proton coupled one electron oxidation of [RuIVRuIII(µ-CO3)4(OH)(CO3)]4− to [RuIVRuIV(µ-CO3)4(O)(CO3)]4− (Equation (23)). After the formation of [RuIVRuIV(µ-CO3)4(O)(CO3)]4−, the axial carbonate decomposes to form [RuIVRuIV(µ-CO3)4(O)2]4− (Equation (24)). Equations (23) and (24) are based on the DFT calculations, see below.
[RuIVRuIII(µ-CO3)4(OH)(CO3)]4− → [RuIVRuIV(µ-CO3)4(O)(CO3)]4− + H+ + e
[RuIVRuIV(µ-CO3)4(O)(CO3)]4− → [RuIVRuIV(µ-CO3)4(O)2]4− + CO2
Since the catalytic process starts at this stage, the current measured represents a process involving many electrons. The dependence of Epa of the third process on the pH is plotted in Figure 9. The slope of the line in Figure 9 is −84 mV·pH−1, thus pointing out that this is a PCET process.

2.2.4. The Fourth Wave, Catalytic Oxidation

The catalytic oxidation of water/HCO3 occurs at potentials ≥ 1.5 V and proceeds via reactions (25)–(31) as derived by DFT, see below:
Two site mechanism:
[RuIVRuIV(µ-CO3)4(O)2]4− + 2H2O → [RuIIIRuIII(µ-CO3)4(OOH)2]4− + 2H+ + 2e
[RuIIIRuIII(µ-CO3)4(OOH)2]4− → [RuIVRuIV(µ-CO3)4(OO)2]4− + 2H+ + 2e
[RuIVRuIV(µ-CO3)4(OO)2]4− + H2O → [RuIIIRuII(µ-CO3)4(OH)]4− + 2O2 + H+ + e
Single site mechanism:
[RuIVRuIV(µ-CO3)4(O)(CO3)]4− + H2O → [RuIVRuII(µ-CO3)4(O)]4− + HCO4 + H+
[RuIVRuII(µ-CO3)4(O)]4− + H2O → [RuIIIRuII(µ-CO3)4(OOH)]4− + H+ + e
[RuIIIRuII(µ-CO3)4(OOH)]4− → [RuIVRuII(µ-CO3)4(OO)]4− + H+ + e
[RuIVRuII(µ-CO3)4(OO)]4− + H2O → [RuIIIRuII(µ-CO3)4(OH)]4− + O2 + H+ + e
The catalytic process is homogeneous as the catalytic peak current increases linearly with the catalyst concentration (Figure 10 and Figure S10).
The catalytic peak/plateau current also depends linearly on [HCO3] (Figure 11 and Figure S11).
Involvement of protons in the catalytic process is demonstrated from the cathodic shift of the onset peak potential (Epo) with the increase in pH with a slope of −48 mV·pH−1 (Figure 12 and Figure S12). Discussing the first oxidation step in the presence of bicarbonate, it was suggested that after one electron oxidation, the species formed is [RuIIIRuIII(µ-CO3)4(OH)2]4− (Equation (22)). The redox behavior in the region 0 to 1.1 to −0.9 V (Figure S13) was studied where in the reverse scan three reduction peaks at 0.40, 0.10, and −0.55 V vs. NHE are observed due to the following processes (Equations (32)–(34)). On the other hand, in the same solution, when an identical scan is performed in the region 0 to −0.9, only one prominent reversible reduction peak is observed at −0.75 V vs. NHE (Figure S14) that is due to the reduction of [RuIIIRuII(µ-CO3)4(OH)]4− to [RuIIRuII(µ-CO3)4]6−. The other peaks are not observed, hence when the scan is first performed up to 1.1 V, then the first hydroxide gets coordinated on the oxidized species and the reduction peaks of various species can be seen. It is important to note that all the processes observed are diffusion controlled and no heterogeneous processes are taking place (Figures S15–S17). The last reduction peak is further studied (Figure S18) and the diffusion coefficient, DR is found to be 2.64 × 10−6 cm2·s−1 and the rate constant of electron transfer, k0 to be (3.12 ± 0.30) × 10−3 cm s−1 (Table S2).
[RuIVRuIII(µ-CO3)4(OH)(CO3)]4− + H2O + e → [RuIIIRuIII(µ-CO3)4(OH)2]4− + HCO3
[RuIIIRuIII(µ-CO3)4(OH)2]4− + e → [RuIIIRuII(µ-CO3)4(OH)]4− + OH
[RuIIIRuII(µ-CO3)4(OH)]4− + e → [RuIIRuII(µ-CO3)4]6− + OH
In the absorption spectra of the [RuIIIRuII(µ-CO3)4]3− complex, two peaks are observed at λ = 337 nm (ε = 336 dm3·mol−1·cm−1) and 412 nm (ε = 834 dm3·mol−1·cm−1) (Figure S19). The first peak is due to the transition from the b2g to the b1u orbital, i.e., δRu-Ru → δ*Ru-Ru and the second one is attributed to the transition from b2g to a1g and a2u, i.e., δRu-Ru → σn/σʹn [62]. For better understanding of the electronic transitions, a molecular orbital picture of the metal-metal bonding is given in Figure S20 [60,62]. A very weak signal is observed at ≈700 nm due to the ligand to metal (RuIII) charge transfer transition (MLCT) [63]. However, when the spectrum is recorded in bicarbonate media, no spectral changes are observed. Thus bicarbonate/carbonates are not bound to the axial position of the Ru atoms in this medium. Further, to check the stability of the complex, WOC long term chronoamperometry was performed at 1.6 V to follow the change in the current. Interestingly, the current does not change/decrease even after 15 h of chronoamperometry (Figure S21). However, in the electronic spectrum, the intensity of the peak at 708 nm is increased (Figure S22) after the chronoamperometry. This is because both the metals are present in a higher oxidation state. Moreover, this peak at 708 nm can also be observed after adding H2O2 to a solution of the complex in bicarbonate medium (Figure S23). This indicates the formation of Ru=O species [64]. Further, activity has been checked by recording 250 voltametric scans successively, and no change in the current was observed (Figure S24). The homogeneous nature of the catalysis has been confirmed by taking scanning electron microscope images before and after the chronoamperometry; no significant precipitate formation on the electrode surface has been observed (Figure S25), and in the energy dispersive X-ray analysis (EDAX), no Ru content has been found.
The Tafel plot obtained by plotting the overpotential (η) vs. log(current density) (Figure S26) gives a slope of 229 mV·dec−1. For water oxidation catalysis, this usually falls in the range 40–200 mV·dec−1 [65], and with the heterogeneous catalysts, NiCo2O4−nanosheets [66], the maximum value of 393 mV·dec−1 has also been reported. A large value of the Tafel slope means that the rate determining step is neither electrochemical nor chemical and is affected by the experimental conditions [67].
Some researcher describe this fact as the involvement of non-homogeneity and local defects due to bubble formation [68,69]. However, in neutral solutions containing no bicarbonate, the Tafel slope is 185 mV·dec−1 (Figure S27). The current of an electrocatalytic process is given by Equation (35), where nc is the number of electrons involved in the catalytic process, the area A of the electrode in cm2, [C] is the bulk concentration of the oxidized species in mol·cm−3, DO is the diffusion coefficient of the oxidized species in cm2·s−1, kcat is the catalytic rate constant in s−1, and F is Faraday constant. Thus, in an electrocatalytic process, the catalytic current (ic) is independent of the scan rate [70,71,72,73,74]. The relation of ic/id with the inverse of the square root of the scan rate (Equation (36)) is obtained by dividing the Randles-Sevcik equation (Equation (20)) by Equation (35).
i c =   n c F A [ C ] ( k c a t D Co ) 1 2
i c i d = 0.359   n c a t n d 3 / 2 k c a t α v
where nd is the number of electrons involved in the diffusion-controlled process, which is 3 here for the RuIVRuIV/RuIIIRuII couple. The scan rate dependent voltammograms are given in Figure S28. The catalytic rate constant, kcat = 1.48 s−1, is obtained from the slope of the plot of ic/id vs. v−1/2 (Figure 13). When the catalysis in neutral medium is considered, the value is 0.10 s−1 (Figures S29 and S30). Closely related values of 1.00 and 2.23 s−1 have been reported by Liobet and co-workers in out/in-[Ru(HL)(trpy)(H2O)]2+ (HL = 1H-pyrazole-3-carboxylic acid, 5-(2-pyridinil)-, ethyl ester). However, there are reports of the ruthenium complex, [Ru(bda)(isoq)2] (H2bda = 2,2′-bipyridine-6,6′-dicarboxylic acid; isoq = isoquinoline) [75] where the kcat is 300 s−1.

2.3. Theoretical Analysis of the Mechanism

To track the water oxidation route in the presence of the rigid bimetallic ruthenium [RuIIIRuII(µ-CO3)4]3− (I0) complex and to better understand the plausible nature of the intermediates formed in situ, DFT analysis was performed.
The plausible intermediates formed during the WOC process are schematically shown in Figure 14, and the relevant geometrical parameters are collected in Table 1. In order to validate the reliability of the computational level, the calculation was initiated following the optimization of the complex [RuIIIRuII(µ-CO3)4]3− (I0). Notably, the computed structural features correspond well to the experimental geometric data and support an electronic quartet ground state [59]. Initially, I0 promptly reacts with water to form -OH coordinated I1 species. The high exothermicity (−102.32 kcal/mol) drives the reaction forward. By moving from I0I1, the Ru–Ru bond is slightly elongated, and the positive partial charges at the Ru centers are reduced (Table 1). The relatively less positive partial charge at the RuIII (ρRu1 = 0.762 e) center compared to RuII (ρRu1 = 0.968 e) is attributed to the charge transfer from the OH ligand, coordinated to the RuIII. The next step involves a proton-coupled electron transfer (PCET) event to generate I2, that is associated with the simultaneous change in the formal oxidation state from RuII→RuIII. This process comprises a favorable energy change of −17.17 kcal/mol. In I2, the two RuIII centers have an almost equal charge (ρRu1/Ru2 = 0.960/0.923 e) and –OHs are ligated symmetrically with the ruthenium centers (Ru1-O1 = 2.087 Å and Ru2-O2 = 2.038 Å, Table 1). HCO3 participates in the next reaction and coordinates as CO32− to the RuIV center by replacing the OH ligand with the simultaneous release of a proton and electron. The formation of I3 from I2 is a slightly uphill process (∆G0 = 3.88 kcal/mol, Figure 14) and is easily accessible when the potential is applied. The one-electron oxidation of I3 is coupled with the release of a proton from the OH ligand, leading to the formation of a RuIV=O species I4 (Figure 14). In I4, the charge discrepancies of the two RuIV centers are due to different extent of charge transfer of the coordinating oxo and CO32− ligands (ρRu1/Ru2 = 1.752/1.969 e). Thereafter, two possible reaction channels can contribute, originating from I4. A single site mechanism, where water is involved in the percarbonate formation reaction and RuIV is reduced to RuII (I5). This finally releases O2 through an exergonic process to re-evolve complex I1 for the next cycle. It is important to note that the partial charges between two RuIV centers in I8 (ρRu1/Ru2 = 1.002/1.002 e) and I10 (ρRu1/Ru2 = 0.788/0.785 e) are symmetrically distributed. The slightly different charge distribution between two RuIII centers is, however, noticed in I9 (ρRu1/Ru2 = 0.394/0.462 e), can be explained by comparing the asymmetrical coordination of the -OOH ligands with the ruthenium centers (Ru1-O1 = 2.043 Å vs. Ru2-O2 = 2.135 Å, Table 1).
Comparing the two different potential pathways, it is clear that the transformation of I4I5 associated with percarbonate formation is more exothermic (∆G0 = −111.51 kcal/mol) than the alternative CO2 dissociation process (i.e., I4I8, ∆G0 = −0.61 kcal/mol), which directs the reaction to follow the single-site mechanism. Moreover, frontier molecular orbital analysis of I5 shows that the lowest unoccupied molecular orbital (α-LUMO) has a significant contribution of the RuIV=O π* interaction, in which the weights of the p(O) and dxz(Ru) atomic orbitals are 64% and 23%, respectively (Figure 15). A similar orbital interaction is also noticed for I8. The atomic orbital contributions correspond to the RuIV=O π* interaction is lower in I8 (p(O) = 10% and dxz(Ru) = 36%), indicating a relatively lower propensity towards nucleophilic water attack [72]. These results agree with previous theoretical findings that a single ruthenium center is mainly involved in the water oxidation process [69].
Moreover, to simulate the experimentally observed ultraviolet-visible (UV-Vis) spectra, we performed time dependent DFT (TD-DFT) calculations. Our results calculated [RuIIIRuIII(µ-CO3)4(OH)2]4− (I1) species revealed a characteristic peak at 426.9 nm that closely approximates the experimentally observed peak at 412 nm.

3. Materials and Methods

The Synthesis of Na3[Ru2(µ-CO3)4], details of the materials used and their sources, the instrumental specification, measurements, and the pretreatment/polishing of the glassy-carbon (GC) electrode (ALS the electrochemical company, Tokyo, Japan) are outlined in the Supplementary Material. The methods for the calculation of diffusion coefficients and electron transfer rate constants are also given in the Supplementary Material.

3.1. Electrochemistry Methods

The electrochemical results were obtained with an EmStat3 instrument (PalmSens, Compact Electrochemical Interfaces, Randhoeve 221, 3995 GA Houten, The Netherlands). The experiments were carried out using a three-electrode setup; glassy-carbon working electrode, diameter 3.0 mm, Ag/AgCl reference electrode, and a Pt counter electrode under N2 atmosphere. Three electrochemical methods: Cyclic voltammetry, square wave voltammetry, and chrono-amperometry were used. All potentials given in this paper, if not specifically stated, are vs. normal hydrogen electrode (NHE) using the conversion: ENHE = EAg/AgCl + 0.198.

3.2. Computational Details

A hybrid functional, Becke’s three-parameter Lee-Yang-Parr (B3LYP) [76,77], was used for geometry optimization using the density functional theory (DFT) method implemented in the Gaussian16 [78] quantum chemistry software (C.01). The Pople’s basis set 6-311+G(d,p) is utilized for all the non-metal atoms, whereas the well-established Stuttgart/Dresden (SDD) basis set with the effective core potential (ECP) was exploited for Ru. In addition, the dispersion effect was imposed using the Grimme D3 correction [79] with Becke-Johnson damping (BJ) during geometry minimization. The vibration frequency analyses were performed at the same theoretical level to ensure the real minima (Nimg = 0) and to obtain the thermodynamic energy corrections. The hydration effect was considered by a self-consistent reaction field (SCRF) approach using Truhler’s SMD model [80] with default parameters for water. In solution, most of the species were defined by 1 (M) standard state, and 55.5 (M) was considered for water. Therefore, for other concentrations (C), additional corrections were made according to the following equation: RTIn(C). The exact calculation of a proton free energy in solution is not straight forward, and thus, we adopted a value of −272.20 kcal mol−1. Natural bond orbital (NBO) analysis implemented in Gaussian16 is used to calculate partial charges and Wiberg indices, which are a measure of bond orders. In order to compare the excitation energies obtained from the experimental UV–vis spectra, TD-DFT calculation was performed utilizing long-range corrected hybrid CAM-B3LYP functional [81]. The 3D images of the optimized structures were captured using CYLview20 visualization software. Unless explicitly stated, all reported energies are the Gibbs free energies in kcal mol−1.

4. Conclusions

The results point out that Na3[RuIIIRuII(µ-CO3)4] acts as an efficient water oxidation catalyst in neutral medium and in the presence of bicarbonate/carbonate. In neutral solutions, the first oxidation peak can be seen, but the other intermediate oxidation steps are not seen before the electrocatalytic water oxidation step. On the other hand, in the bicarbonate medium before water oxidation the formation of other intermediate species are observed in cyclic voltammetry. It is found that all the electrochemical processes involved are proton coupled electron transfers (PCET). The catalytic peak/plateau current depends linearly on the catalyst and on the bicarbonate concentrations. Thus, it is clear that bicarbonate is involved in the process of oxidation. Furthermore, after the formation of [RuIVRuIII(µ-CO3)4(OH)(CO3)]2− by three electron oxidations when the reverse scan is performed, all the other steps can be seen in the presence of bicarbonate, which cannot be observed in the absence of bicarbonate. This catalyst has the advantage that it contains no organic ligand and is therefore stable during the catalytic cycles.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/11/2/281/s1. Table S1: The values of anodic peak (Epa (V)) and cathodic (Epc (V)) peak potential, their difference, ΔEp (V); Nicholson parameter, ψ and rate constant of electron transfer, k0 (cm s−1) at different scan rate (V s−1) for the redox couple RuIIIRuIII/RuIIIRuII in neutral medium, Table S2: The values of Epa (V), Epc (V), ΔEp (V), ψ and k0 (cm s−1) at different scan rate (V s−1) for the redox couple RuIVRuV/ RuIVRuIV in bicarbonate medium, Table S3: The values of Epa (V), Epc (V), ΔEp (V), ψ and k0 (cm s−1) at different scan rate (V s−1) for the redox couple for the redox couple RuIIRuIII/ RuIIRuII in bicarbonate medium, Figure S1: The crystal structure of Na3[Ru(µ-CO3)4] showing the axial coordination of carbonate ligand from another complex. CCDC No. 1200939, Figure S2: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO (pH 7.0) and 1 mM K4[Fe(CN)6] and 0.20 M NaClO at a scan rate of 50 mV∙s−1, Figure S3: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 (pH 7.0) at various scan rates, Figure S4: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 solution at different pHs with a scan rate of 50 mV∙s−1 highlighting the first redox couple, Figure S5: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 solution at different pHs with a scan rate of 50 mV∙s−1, Figure S6: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] and 1.0 mM K4[Fe(CN)6] in 0.10 M NaHCO3 (pH 8.3) at a scan rate of 50 mV∙s−1, Figure S7. CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 solution at different pHs with a scan rate of 50 mV∙s−1, Figure S8: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) at various scan rates highlighting the first two redox processes, Figure S9: CVs of increasing concentrations of NaHCO3 (pH 8.3) in 1.0 mM Na3[Ru2(µ-CO3)4] with a scan rate of 50 mV∙s−1 highlighting the RuIIIRuIII/RuIIIRuII redox couple, Figure S10: CVs of increasing concentrations of Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) with a scan rate of 50 mV.s−1, Figure S11: CVs of increasing concentrations of NaHCO3 (pH 8.3) in 1.0 mM Na3[Ru2(µ-CO3)4] with a scan rate of 50 mV∙s−1, Figure S12. The CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.19 M NaHCO3 at various pHs with a scan rate of 50 mV s1, Figure S13: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) at various scan rates, Figure S14: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) at various scan rates, Figure S15: id at Epc of 0.11 V vs. v1/2 in 1.0 mM Na3[Ru2(µ−CO3)4] and 0.10 M NaHCO3 (pH 8.3), Figure S16: id at Epc of -0.50 V vs. v1/2 in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 (pH 8.3), Figure S17: id at Epc of −0.75 V vs. v1/2 in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 (pH 8.3), Figure S18: CVs of 5.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) highlighting the RuIIIRuII/RuIIRuII couple, Figure S19: Absorption spectra of 1.0 mM Na3[Ru2(µ-CO3)4] in presence and absence of 0.10 M NaHCO3, Figure S20: A qualitative MO diagram of the metal-metal bonding for an M2X8 (M = transition metal and X is halide) species of symmetry D4h. The electron distribution shown is that for [Ru2(µ-CO3)4]3+, Figure S21: The chronoamperometry (CA) of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 at pH 8.3 for 15 h at a potential of 1.6 V vs. NHE, Figure S22: Absorption spectra before and after chronoamperometry of a solution containing 1.0 mM Na3[Ru2(µ-CO3)4], 0.10 M NaHCO3 at 1.6 V vs. NHE for 15 h, Figure S23: Comparison of the electronic absorption spectra of 1.0 mM Na3[Ru2(µ-CO3)4], 0.10 M NaHCO3 in presence and absence of NaHCO3, Figure S24: The SEM image of the GC electrode before and after chronoamperometry, Figure S25: Successive CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH 8.3) at a scan rate of 50 mV s−1, Figure S26: Tafel plot in presence of bicarbonate. LSV is recorded in 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 at pH 8.3 at a scan rate of 50 mV.s−1, Figure S27: Tafel plot under neutral condition. LSV is recorded in a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 at pH 7.0 at a scan rate of 50 mV.s−1, Figure S28: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 at pH 8.3 at various scan rates, Figure S29: CVs of 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 at pH 7.0 at various scan rates, Figure S30: A Plot of ic/id vs. v−1/2 for a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.20 M NaClO4 (pH 7.0). The ic is taken at 1.56 V.

Author Contributions

Conceptualization, S.G.P. and D.M.; methodology, S.G.P., T.M., K.S., D.M., H.K., and A.M.; software, H.K.; validation, S.G.P., D.M., and H.K.; investigation, S.G.P., T.M., K.S., D.M., H.K., and A.M.; resources, D.M.; writing—original draft preparation, S.G.P. and D.M.; writing—review and editing, S.G.P., D.M., T.M., H.K., A.M.; visualization, S.G.P. and T.M.; supervision, D.M. and H.K.; project administration, D.M.; funding acquisition, D.M., and A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by PAZY FOUNDATION, grant number RA 1700000337.

Data Availability Statement

Data available on request. Some data can also be found in the Supplementary Material.

Acknowledgments

We are indebted to the Pazy foundation for financial support that enabled this study. SGP thanks Ariel University and Dan Meyerstein for post-doctoral fellowship.

Conflicts of Interest

The authors declare no conflict of interest

References

  1. Shatskiy, A.; Kärkäs, M.D.; Åkermark, B. The Art of Splitting Water: Storing Energy in a Readily Available and Convenient Form. Eur. J. Inorg. Chem. 2019, 2019, 2020–2024. [Google Scholar] [CrossRef] [Green Version]
  2. Dresselhaus, M.S.; Thomas, I.L. Alternative energy technologies. Nature 2001, 414, 332–337. [Google Scholar] [CrossRef] [PubMed]
  3. Ort, D.R.; Yocum, C.F.; Heichel, I.F. (Eds.) Oxygenic photosynthesis: The light reactions. In Advances in Photosynthesis and Respiration; Springer: Dordrecht, The Netherlands, 1996; Volume 4, ISBN 978-0-7923-3683-9. [Google Scholar]
  4. Meyer, T.J.; Huynh, M.H.V.; Thorp, H.H. The possible role of proton-coupled electron transfer (PCET) in water oxidation by photosystem II. Angew. Chemie Int. Ed. 2007, 46, 5284–5304. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, Y.; Kraft, M.; Xu, R. Metal-free carbonaceous electrocatalysts and photocatalysts for water splitting. Chem. Soc. Rev. 2016, 45, 3039–3052. [Google Scholar] [CrossRef]
  6. Tian, T.; Zheng, M.; Lin, J.; Meng, X.; Ding, Y. Amorphous Ni–Fe double hydroxide hollow nanocubes enriched with oxygen vacancies as efficient electrocatalytic water oxidation catalysts. Chem. Commun. 2019, 55, 1044–1047. [Google Scholar] [CrossRef] [PubMed]
  7. Fukuzumi, S.; Lee, Y.M.; Nam, W. Kinetics and mechanisms of catalytic water oxidation. Dalt. Trans. 2019, 48, 779–798. [Google Scholar] [CrossRef]
  8. Lee, Y.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao-Horn, Y. Synthesis and Activities of Rutile IrO2 and RuO2 Nanoparticles for Oxygen Evolution in Acid and Alkaline Solutions. J. Phys. Chem. Lett. 2012, 3, 399–404. [Google Scholar] [CrossRef] [PubMed]
  9. Peng, L.; Shah, S.S.A.; Wei, Z. Recent developments in metal phosphide and sulfide electrocatalysts for oxygen evolution reaction. Chin. J. Catal. 2018, 39, 1575–1593. [Google Scholar] [CrossRef]
  10. Hunter, B.M.; Gray, H.B.; Müller, A.M. Earth-Abundant Heterogeneous Water Oxidation Catalysts. Chem. Rev. 2016, 116, 14120–14136. [Google Scholar] [CrossRef] [PubMed]
  11. Wasylenko, D.J.; Palmer, R.D.; Berlinguette, C.P. Homogeneous water oxidation catalysts containing a single metal site. Chem. Commun. 2013, 49, 218–227. [Google Scholar] [CrossRef]
  12. Parent, A.R.; Sakai, K. Progress in base-metal water oxidation catalysis. ChemSusChem 2014, 7, 2070–2080. [Google Scholar] [CrossRef]
  13. Indra, A.; Menezes, P.W.; Driess, M. Uncovering structure-activity relationships in manganese-oxide-based heterogeneous catalysts for efficient water oxidation. ChemSusChem 2015, 8, 776–785. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, T.; Zhang, B.; Sun, L. Iron-Based Molecular Water Oxidation Catalysts: Abundant, Cheap, and Promising. Chem. Asian J. 2019, 14, 31–43. [Google Scholar] [CrossRef]
  15. Kärkäs, M.D.; Åkermark, B. Water oxidation using earth-abundant transition metal catalysts: Opportunities and challenges. Dalt. Trans. 2016, 45, 14421–14461. [Google Scholar] [CrossRef] [Green Version]
  16. Singh, A.; Spiccia, L. Water oxidation catalysts based on abundant 1st row transition metals. Coord. Chem. Rev. 2013, 257, 2607. [Google Scholar] [CrossRef]
  17. Risch, M.; Klingan, K.; Zaharieva, I.; Dau, H. Water Oxidation by Co-Based Oxides with Molecular Properties. Mol. Water Oxid. Catal. Key Top. New Sustain. Energy Convers. Schemes 2014, 9781118413, 163–185. [Google Scholar] [CrossRef]
  18. Van Oversteeg, C.H.M.; Doan, H.Q.; De Groot, F.M.F.; Cuk, T. In situ X-ray absorption spectroscopy of transition metal based water oxidation catalysts. Chem. Soc. Rev. 2017, 46, 102–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Basu, O.; Mukhopadhyay, S.; Das, S.K. Cobalt based functional inorganic materials: Electrocatalytic water oxidation. J. Chem. Sci. 2018, 130, 1–15. [Google Scholar] [CrossRef] [Green Version]
  20. Patra, S.G.; Illés, E.; Mizrahi, A.; Meyerstein, D. Cobalt Carbonate as an Electrocatalyst for Water Oxidation. Chem. A Eur. J. 2020, 26, 711–720. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, J.W.; Liu, W.J.; Zhong, D.C.; Lu, T.B. Nickel complexes as molecular catalysts for water splitting and CO2 reduction. Coord. Chem. Rev. 2019, 378, 237–261. [Google Scholar] [CrossRef]
  22. Ariela, B.; Yaniv, W.; Dror, S.; Haya, K.; Yael, A.; Eric, M.; Dan, M. The role of carbonate in electro-catalytic water oxidation by using Ni(1,4,8,11-tetraazacyclotetradecane) 2+. Dalt. Trans. 2017, 46, 10774–10779. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, S.; Lei, Y.J.; Xin, Z.J.; Lu, Y.B.; Wang, H.Y. Water splitting based on homogeneous copper molecular catalysts. J. Photochem. Photobiol. A Chem. 2018, 355, 141–151. [Google Scholar] [CrossRef]
  24. Shaffer, D.W.; Xie, Y.; Concepcion, J.J. O-O bond formation in ruthenium-catalyzed water oxidation: Single-site nucleophilic attack: Vs. O-O radical coupling. Chem. Soc. Rev. 2017, 46, 6170–6193. [Google Scholar] [CrossRef] [PubMed]
  25. Corbucci, I.; Macchioni, A.; Albrecht, M. Iridium Complexes in Water Oxidation Catalysis. Iridium(III) Optoelectron. Photonics Appl. 2017, 617–654. [Google Scholar] [CrossRef]
  26. Matheu, R.; Ertem, M.Z.; Gimbert-Suriñach, C.; Sala, X.; Llobet, A. Seven Coordinated Molecular Ruthenium-Water Oxidation Catalysts: A Coordination Chemistry Journey. Chem. Rev. 2019, 119, 3453–3471. [Google Scholar] [CrossRef]
  27. Blakemore, J.D.; Crabtree, R.H.; Brudvig, G.W. Molecular Catalysts for Water Oxidation. Chem. Rev. 2015, 115, 12974–13005. [Google Scholar] [CrossRef] [PubMed]
  28. Gersten, S.W.; Samuels, G.J.; Meyer, T.J. Catalytic oxidation of water by an oxo-bridged ruthenium dimer. J. Am. Chem. Soc. 1982, 104, 4029–4030. [Google Scholar] [CrossRef]
  29. Zhang, B.; Sun, L. Artificial photosynthesis: Opportunities and challenges of molecular catalysts. Chem. Soc. Rev. 2019, 48, 2216–2264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Stull, J.A.; Stich, T.A.; Hurst, J.K.; Britt, R.D. Electron paramagnetic resonance analysis of a transient species formed during water oxidation catalyzed by the complex ion [(bpy)2Ru(OH2)]2O4+. Inorg. Chem. 2013, 52, 4578–4586. [Google Scholar] [CrossRef]
  31. Yamada, H.; Siems, W.F.; Koike, T.; Hurst, J.K. Mechanisms of water oxidation catalyzed by the cis,cis-[(bpy)2Ru(OH2)]2O4+ ion. J. Am. Chem. Soc. 2004, 126, 9786–9795. [Google Scholar] [CrossRef]
  32. Moonshiram, D.; Jurss, J.W.; Concepcion, J.J.; Zakharova, T.; Alperovich, I.; Meyer, T.J.; Pushkar, Y. Structure and electronic configurations of the intermediates of water oxidation in blue ruthenium dimer catalysis. J. Am. Chem. Soc. 2012, 134, 4625–4636. [Google Scholar] [CrossRef] [PubMed]
  33. Yamada, H.; Hurst, J.K. Resonance raman, optical spectroscopic, and EPR characterization of the higher oxidation states of the water oxidation catalyst, cis,cis-[(bpy)2Ru(OH2)]2O4+. J. Am. Chem. Soc. 2000, 122, 5303–5311. [Google Scholar] [CrossRef]
  34. Moonshiram, D.; Alperovich, I.; Concepcion, J.J.; Meyer, T.J.; Pushkar, Y. Experimental demonstration of radicaloid character in a RuV=O intermediate in catalytic water oxidation. Proc. Natl. Acad. Sci. USA 2013, 110, 3765–3770. [Google Scholar] [CrossRef] [Green Version]
  35. Cape, J.L.; Hurst, J.K. Detection and mechanistic relevance of transient ligand radicals formed during [Ru(bpy)2(OH2)]2O4+-catalyzed water oxidation. J. Am. Chem. Soc. 2008, 130, 827–829. [Google Scholar] [CrossRef]
  36. Najafpour, M.M.; Renger, G.; Hołyńska, M.; Moghaddam, A.N.; Aro, E.M.; Carpentier, R.; Nishihara, H.; Eaton-Rye, J.J.; Shen, J.R.; Allakhverdiev, S.I. Manganese Compounds as Water-Oxidizing Catalysts: From the Natural Water-Oxidizing Complex to Nanosized Manganese Oxide Structures. Chem. Rev. 2016, 116, 2886–2936. [Google Scholar] [CrossRef] [PubMed]
  37. Moyer, B.A.; Meyer, T.J. Properties of the oxo/aqua system (bpy)2(py)RuO2+/(bpy)2(py)Ru(OH2)2+. Inorg. Chem. 1981, 20, 436–444. [Google Scholar] [CrossRef]
  38. Moyer, B.A.; Meyer, T.J. Oxobis(2,2’-bipyridine)pyridineruthenium(IV) ion, [(bpy)2(py)Ru:O]2+. J. Am. Chem. Soc. 1978, 100, 3601–3603. [Google Scholar] [CrossRef]
  39. Tong, L.; Thummel, R.P. Mononuclear ruthenium polypyridine complexes that catalyze water oxidation. Chem. Sci. 2016, 7, 6591–6603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Gagliardi, C.J.; Vannucci, A.K.; Concepcion, J.J.; Chen, Z.; Meyer, T.J. The role of proton coupled electron transfer in water oxidation. Energy Environ. Sci. 2012, 5, 7704. [Google Scholar] [CrossRef]
  41. Jenson, D.L.; Barry, B.A. Proton-coupled electron transfer in photosystem II: Proton inventory of a redox active tyrosine. J. Am. Chem. Soc. 2009, 131, 10567–10573. [Google Scholar] [CrossRef] [Green Version]
  42. Mizrahi, A.; Meyerstein, D. Plausible roles of carbonate in catalytic water oxidation. In Advances in Inorganic Chemistry; Elsevier: Amsterdam, The Netherlands, 2019; pp. 343–360. [Google Scholar]
  43. Patra, S.G.; Mizrahi, A.; Meyerstein, D. The Role of Carbonate in Catalytic Oxidations. Acc. Chem. Res. 2020, 53, 2189–2200. [Google Scholar] [CrossRef]
  44. Steenken, S.; Neta, P.; Stanbury, D.M.; Armstrong, D.A.; Ruscic, B.; Koppenol, W.H.; Huie, R.E.; Merényi, G.; Wardman, P.; Lymar, S.V. Standard electrode potentials involving radicals in aqueous solution: Inorganic radicals (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1139–1150. [Google Scholar] [CrossRef]
  45. Zilberg, S.; Mizrahi, A.; Meyerstein, D.; Kornweitz, H. Carbonate and carbonate anion radicals in aqueous solutions exist as CO3(H2O)62− and CO3(H2O)6 respectively: The crucial role of the inner hydration sphere of anions in explaining their properties. Phys. Chem. Chem. Phys. 2018, 20, 9429–9435. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, J.; Ji, L.; Chen, Z. In Situ Rapid Formation of a Nickel–Iron-Based Electrocatalyst for Water Oxidation. ACS Catal. 2016, 6, 6987–6992. [Google Scholar] [CrossRef]
  47. Khorobrykh, A.; Dasgupta, J.; Kolling, D.R.J.; Terentyev, V.; Klimov, V.V.; Dismukes, G.C. Evolutionary Origins of the Photosynthetic Water Oxidation Cluster: Bicarbonate Permits Mn2+ Photo-oxidation by Anoxygenic Bacterial Reaction Centers. ChemBioChem 2013, 14, 1725–1731. [Google Scholar] [CrossRef]
  48. Hobart, D.E.; Samhoun, K.; Young, J.P.; Norvell, V.E.; Mamantov, G.; Peterson, J.R. Stabilization of praseodymium(IV) and terbium(IV) in aqueous carbonate solution. Inorg. Nucl. Chem. Lett. 1980, 16, 321–328. [Google Scholar] [CrossRef]
  49. Koroidov, S.; Shevela, D.; Shutova, T.; Samuelsson, G.; Messinger, J. Mobile hydrogen carbonate acts as proton acceptor in photosynthetic water oxidation. Proc. Natl. Acad. Sci. USA 2014, 111, 6299–6304. [Google Scholar] [CrossRef] [Green Version]
  50. Shevela, D.; Nöring, B.; Koroidov, S.; Shutova, T.; Samuelsson, G.; Messinger, J. Efficiency of photosynthetic water oxidation at ambient and depleted levels of inorganic carbon. Photosynth. Res. 2013, 117, 401–412. [Google Scholar] [CrossRef]
  51. Chen, Z.; Meyer, T.J. Copper(II) Catalysis of Water Oxidation. Angew. Chemie Int. Ed. 2013, 52, 700–703. [Google Scholar] [CrossRef]
  52. Winikoff, S.G.; Cramer, C.J. Mechanistic analysis of water oxidation catalyzed by mononuclear copper in aqueous bicarbonate solutions. Catal. Sci. Technol. 2014, 4, 2484–2489. [Google Scholar] [CrossRef]
  53. Mizrahi, A.; Maimon, E.; Cohen, H.; Kornweitz, H.; Zilbermann, I.; Meyerstein, D. Mechanistic Studies on the Role of [CuII(CO3)n]2−2n as a Water Oxidation Catalyst: Carbonate as a Non-Innocent Ligand. Chem. A Eur. J. 2018, 24, 1088–1096. [Google Scholar] [CrossRef]
  54. Chen, F.; Wang, N.; Lei, H.; Guo, D.; Liu, H.; Zhang, Z.; Zhang, W.; Lai, W.; Cao, R. Electrocatalytic Water Oxidation by a Water-Soluble Copper(II) Complex with a Copper-Bound Carbonate Group Acting as a Potential Proton Shuttle. Inorg. Chem. 2017, 56, 13368–13375. [Google Scholar] [CrossRef]
  55. Kuttassery, F.; Sebastian, A.; Mathew, S.; Tachibana, H.; Inoue, H. Promotive Effect of Bicarbonate Ion on Two-Electron Water Oxidation to Form H2O2 Catalyzed by Aluminum Porphyrins. ChemSusChem 2019, 12, 1939–1948. [Google Scholar] [CrossRef]
  56. Haygarth, K.S.; Marin, T.W.; Janik, I.; Kanjana, K.; Stanisky, C.M.; Bartels, D.M. Carbonate radical formation in radiolysis of sodium carbonate and bicarbonate solutions up to 250 °C and the mechanism of its second order decay. J. Phys. Chem. A 2010, 114, 2142–2150. [Google Scholar] [CrossRef]
  57. Lindsay, A.J.; Wilkinson, G.; Motevalli, M.; Hursthouse, M.B. Reactions of tetra-µ-carboxylato-diruthenium(II,II) compounds. X-Ray crystal structures of Ru2(µ-O2CCF3)4(thf)2, Ru2(µ-O2CR)4. J. Chem. Soc. Dalt. Trans. 1987, 2723–2736. [Google Scholar] [CrossRef]
  58. Lindsay, A.J.; Motevalli, M.; Hursthouse, M.B.; Wilkinson, G. The synthesis and structure of the unusual polymeric compound {Na3[Ru2(µ-O2CO)4]·6H2O}n. J. Chem. Soc. Chem. Commun. 1986, 2, 433–434. [Google Scholar] [CrossRef]
  59. Cotton, F.A.; Labella, L.; Shang, M. Further study of tetracarbonato diruthenium(II,III) compounds. Inorg. Chem. 1992, 31, 2385–2389. [Google Scholar] [CrossRef]
  60. Cotton, F.A.; Murillo, C.A.; Walton, R.A. Multiple Bonds between Metal Atoms; Cotton, F.A., Murillo, C.A., Walton, R.A., Eds.; Springer: New York, NY, USA, 2005; ISBN 0-387-25084-0. [Google Scholar]
  61. Nicholson, R.S. Theory and Application of Cyclic Voltammetry for Measurement of Electrode Reaction Kinetics. Anal. Chem. 1965, 37, 1351–1355. [Google Scholar] [CrossRef]
  62. Bennett, M.J.; Caulton, K.G.; Cotton, F.A. Structure of tetra-n-butyratodiruthenium chloride, a compound with a strong metal-metal bond. Inorg. Chem. 1969, 8, 1–6. [Google Scholar] [CrossRef]
  63. Guadalupe Hernández, J.; Huerta-Aguilar, C.A.; Thangarasu, P.; Höpfl, H. A ruthenium(iii) complex derived from N,N′-bis(salicylidene)ethylenediamine as a chemosensor for the selective recognition of acetate and its interaction with cells for bio-imaging: Experimental and theoretical studies. New J. Chem. 2017, 41, 10815–10827. [Google Scholar] [CrossRef]
  64. Wada, T.; Nishimura, S.; Mochizuki, T.; Ando, T.; Miyazato, Y. Mechanism of Water Oxidation Catalyzed by a Dinuclear Ruthenium Complex Bridged by Anthraquinone. Catalysts 2017, 7, 56. [Google Scholar] [CrossRef]
  65. Bockris, J.O. The Electrocatalysis of Oxygen Evolution on Perovskites. J. Electrochem. Soc. 1984, 131, 290. [Google Scholar] [CrossRef]
  66. Shi, H.; Zhao, G. Water Oxidation on Spinel NiCo2O4 Nanoneedles Anode: Microstructures, Specific Surface Character, and the Enhanced Electrocatalytic Performance. J. Phys. Chem. C 2014, 118, 25939–25946. [Google Scholar] [CrossRef]
  67. Goberna-Ferrón, S.; Peña, B.; Soriano-López, J.; Carbó, J.J.; Zhao, H.; Poblet, J.M.; Dunbar, K.R.; Galán-Mascarós, J.R. A fast metal–metal bonded water oxidation catalyst. J. Catal. 2014, 315, 25–32. [Google Scholar] [CrossRef]
  68. Blakemore, J.D.; Schley, N.D.; Olack, G.W.; Incarvito, C.D.; Brudvig, G.W.; Crabtree, R.H. Anodic deposition of a robust iridium-based water-oxidation catalyst from organometallic precursors. Chem. Sci. 2011, 2, 94–98. [Google Scholar] [CrossRef] [Green Version]
  69. Schley, N.D.; Blakemore, J.D.; Subbaiyan, N.K.; Incarvito, C.D.; D’Souza, F.; Crabtree, R.H.; Brudvig, G.W. Distinguishing Homogeneous from Heterogeneous Catalysis in Electrode-Driven Water Oxidation with Molecular Iridium Complexes. J. Am. Chem. Soc. 2011, 133, 10473–10481. [Google Scholar] [CrossRef]
  70. Wang, D.; Groves, J.T. Efficient water oxidation catalyzed by homogeneous cationic cobalt porphyrins with critical roles for the buffer base. Proc. Natl. Acad. Sci. USA 2013, 110, 15579–15584. [Google Scholar] [CrossRef] [Green Version]
  71. Vannucci, A.K.; Meyer, T.J.; Dares, C.J.; Coggins, M.K.; Zhang, M.-T. Electrocatalytic Water Oxidation by a Monomeric Amidate-Ligated Fe(III)–Aqua Complex. J. Am. Chem. Soc. 2014, 136, 5531–5534. [Google Scholar] [CrossRef]
  72. Koepke, S.J.; Light, K.M.; Vannatta, P.E.; Wiley, K.M.; Kieber-Emmons, M.T. Electrocatalytic Water Oxidation by a Homogeneous Copper Catalyst Disfavors Single-Site Mechanisms. J. Am. Chem. Soc. 2017, 139, 8586–8600. [Google Scholar] [CrossRef]
  73. Wang, D.; Bruner, C.O. Catalytic Water Oxidation by a Bio-inspired Nickel Complex with a Redox-Active Ligand. Inorg. Chem. 2017, 56, 13638–13641. [Google Scholar] [CrossRef] [Green Version]
  74. Du, H.Y.; Chen, S.C.; Su, X.J.; Jiao, L.; Zhang, M.T. Redox-Active Ligand Assisted Multielectron Catalysis: A Case of CoIII Complex as Water Oxidation Catalyst. J. Am. Chem. Soc. 2018, 140, 1557–1565. [Google Scholar] [CrossRef]
  75. Duan, L.; Bozoglian, F.; Mandal, S.; Stewart, B.; Privalov, T.; Llobet, A.; Sun, L. A molecular ruthenium catalyst with water-oxidation activity comparable to that of photosystem II. Nat. Chem. 2012, 4, 418–423. [Google Scholar] [CrossRef]
  76. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  77. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  78. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  79. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef] [PubMed]
  81. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Cyclic voltammograms of a solution containing 0.20 M NaClO4 (pH 7.0) in the presence and absence of 1.0 mM Na3[Ru2(µ-CO3)4] at a scan rate of 50 mV·s−1 under N2 atmosphere.
Figure 1. Cyclic voltammograms of a solution containing 0.20 M NaClO4 (pH 7.0) in the presence and absence of 1.0 mM Na3[Ru2(µ-CO3)4] at a scan rate of 50 mV·s−1 under N2 atmosphere.
Catalysts 11 00281 g001
Figure 2. A plot of the anodic peak current (id) of the RuIIIRuIII/RuIIIRuII peak vs. the square root of the scan rate. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 (pH 7.0) under N2 atmosphere.
Figure 2. A plot of the anodic peak current (id) of the RuIIIRuIII/RuIIIRuII peak vs. the square root of the scan rate. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 (pH 7.0) under N2 atmosphere.
Catalysts 11 00281 g002
Figure 3. A plot of the anodic peak potential (Epa) of the first redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Figure 3. A plot of the anodic peak potential (Epa) of the first redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Catalysts 11 00281 g003
Figure 4. A plot of the water oxidation peak/plateau potential vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Figure 4. A plot of the water oxidation peak/plateau potential vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.20 M NaClO4 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Catalysts 11 00281 g004
Figure 5. Cyclic voltammograms of a solution containing 1.0 mM Na3[Ru2-CO3)4] in the presence of 0.10 M NaHCO3 (pH 8.3, blue line) and 0.20 M NaClO4 (pH 7.0, black line) scan rate, 50 mV s−1 under N2 atmosphere.
Figure 5. Cyclic voltammograms of a solution containing 1.0 mM Na3[Ru2-CO3)4] in the presence of 0.10 M NaHCO3 (pH 8.3, blue line) and 0.20 M NaClO4 (pH 7.0, black line) scan rate, 50 mV s−1 under N2 atmosphere.
Catalysts 11 00281 g005
Figure 6. Square wave voltammogram of a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in the presence of 0.10 M NaHCO3, pH 8.3, under N2 atmosphere. Instrumental settings: Amplitude, 0.010 V; frequency, 15.0 Hz.
Figure 6. Square wave voltammogram of a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in the presence of 0.10 M NaHCO3, pH 8.3, under N2 atmosphere. Instrumental settings: Amplitude, 0.010 V; frequency, 15.0 Hz.
Catalysts 11 00281 g006
Figure 7. The cyclic voltammograms of the redox wave are recorded in various bicarbonate concentrations. However, no significant change in the peak potential/current was found (Figure S9). This is because bicarbonate replaces the hydroxide ion after the redox process and is not involved in the electron transfer step.
Figure 7. The cyclic voltammograms of the redox wave are recorded in various bicarbonate concentrations. However, no significant change in the peak potential/current was found (Figure S9). This is because bicarbonate replaces the hydroxide ion after the redox process and is not involved in the electron transfer step.
Catalysts 11 00281 g007
Figure 8. A plot of the anodic peak potential (Epa) of the RuIVRuIII/RuIIIRuIII redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Figure 8. A plot of the anodic peak potential (Epa) of the RuIVRuIII/RuIIIRuIII redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Catalysts 11 00281 g008
Figure 9. A plot of the anodic peak potential (Epa) of the RuIVRuIV/RuIVRuIII redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Figure 9. A plot of the anodic peak potential (Epa) of the RuIVRuIV/RuIVRuIII redox couple vs. the pH of the medium. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] and 0.10 M NaHCO3 aqueous solutions at different pHs with a scan rate of 50 mV·s−1 under N2 atmosphere.
Catalysts 11 00281 g009
Figure 10. A linear increase of the catalytic peak current at 1.55 V vs. [Na3[Ru2(µ-CO3)4]]. The voltammograms are recorded in 0.10 M NaHCO3 (pH 8.3) under N2 atmosphere with a scan rate of 50 mV·s−1.
Figure 10. A linear increase of the catalytic peak current at 1.55 V vs. [Na3[Ru2(µ-CO3)4]]. The voltammograms are recorded in 0.10 M NaHCO3 (pH 8.3) under N2 atmosphere with a scan rate of 50 mV·s−1.
Catalysts 11 00281 g010
Figure 11. A plot of the catalytic peak/plateau current at 1.55 V vs. [NaHCO3]. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] with increasing concentrations of NaHCO3 (pH 8.3) at a scan rate of 50 mV·s−1.
Figure 11. A plot of the catalytic peak/plateau current at 1.55 V vs. [NaHCO3]. The voltammograms are recorded in 1.0 mM Na3[Ru2(µ-CO3)4] with increasing concentrations of NaHCO3 (pH 8.3) at a scan rate of 50 mV·s−1.
Catalysts 11 00281 g011
Figure 12. Variation of the catalytic water oxidation potential with pH. The voltammograms are recorded in solutions containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 at various pHs under N2 atmosphere, scan rate 50 mV·s−1. The onset potential is taken at a current of 400 µA.
Figure 12. Variation of the catalytic water oxidation potential with pH. The voltammograms are recorded in solutions containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 at various pHs under N2 atmosphere, scan rate 50 mV·s−1. The onset potential is taken at a current of 400 µA.
Catalysts 11 00281 g012
Figure 13. A plot of ic/id vs. v−1/2 for a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH—8.3). The ic is taken at 1.56 V.
Figure 13. A plot of ic/id vs. v−1/2 for a solution containing 1.0 mM Na3[Ru2(µ-CO3)4] in 0.10 M NaHCO3 (pH—8.3). The ic is taken at 1.56 V.
Catalysts 11 00281 g013
Figure 14. Schematic mechanism of the electrocatalytic water oxidation in the presence of [RuIIIRuII(µ-CO3)4]3− (I0). The free energies of the reactions are highlighted in yellow (ΔG, kcal/mol). The grey part represents the double site mechanism and the black one the single site mechanism.
Figure 14. Schematic mechanism of the electrocatalytic water oxidation in the presence of [RuIIIRuII(µ-CO3)4]3− (I0). The free energies of the reactions are highlighted in yellow (ΔG, kcal/mol). The grey part represents the double site mechanism and the black one the single site mechanism.
Catalysts 11 00281 g014
Figure 15. LUMO representations for the species [RuIVRuII(µ-CO3)4(O)]4− (I5) and [RuIVRuIV(µ-CO3)4(O)2]4− (I8).
Figure 15. LUMO representations for the species [RuIVRuII(µ-CO3)4(O)]4− (I5) and [RuIVRuIV(µ-CO3)4(O)2]4− (I8).
Catalysts 11 00281 g015
Table 1. Electronic and structural data for species involved in the water oxidation catalysts (WOC) mechanism.
Table 1. Electronic and structural data for species involved in the water oxidation catalysts (WOC) mechanism.
SpeciesConfigurationSpinq ad bRu1-Ru2d cRu1-O1/Ru2-O2ρ dRu1/Ru2
I0RuII-RuIII
d6-d5
3/2−32.235(2.260) e-1.052/1.074
I1RuII-RuIII
d6-d5
3/2−42.334-/2.0700.968/0.762
I2RuIII-RuIII
d5-d5
1−42.3182.087/2.0380.960/0.923
I3RuIII-RuIV
d5-d4
1/2−42.4582.059/1.9281.160/1.058
I4RuIV-RuIV
d4-d4
0−42.4581.752/1.9691.084/0.991
I5RuIV-RuII
d4-d6
1−42.4221.850/-0.839/1.027
I6RuIII-RuII
d5-d6
3/2−42.3032.023/-0.712/0.979
I7RuIV-RuII
d4-d6
1−42.3472.094/-0.731/1.128
I8RuIV-RuIV
d4-d4
0−42.4441.769/1.7691.002/1.002
I9RuIII-RuIII
d5-d5
1−42.3422.043/2.1350.394/0.462
I10RuIV-RuIV
d4-d4
0−42.5232.080/2.0800.788/0.785
a Overall system charge. b Ru1-Ru2 bond distances in Å. c Ruthenium-oxygen (ligands coordinating during WOC process). d Partial natural bond orbital (NBO) charges. e Value in parenthesis correspond to X-ray bond distance.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patra, S.G.; Mondal, T.; Sathiyan, K.; Mizrahi, A.; Kornweitz, H.; Meyerstein, D. Na3[Ru2(µ-CO3)4] as a Homogeneous Catalyst for Water Oxidation; HCO3 as a Co-Catalyst. Catalysts 2021, 11, 281. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020281

AMA Style

Patra SG, Mondal T, Sathiyan K, Mizrahi A, Kornweitz H, Meyerstein D. Na3[Ru2(µ-CO3)4] as a Homogeneous Catalyst for Water Oxidation; HCO3 as a Co-Catalyst. Catalysts. 2021; 11(2):281. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020281

Chicago/Turabian Style

Patra, Shanti Gopal, Totan Mondal, Krishnamoorthy Sathiyan, Amir Mizrahi, Haya Kornweitz, and Dan Meyerstein. 2021. "Na3[Ru2(µ-CO3)4] as a Homogeneous Catalyst for Water Oxidation; HCO3 as a Co-Catalyst" Catalysts 11, no. 2: 281. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020281

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop