Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 May 17;145(21):11818–11828. doi: 10.1021/jacs.3c03415

Bioinspired Active Site with a Coordination-Adaptive Organosulfonate Ligand for Catalytic Water Oxidation at Neutral pH

Tianqi Liu , Shaoqi Zhan ‡,§, Nannan Shen , Linqin Wang , Zoltán Szabó , Hao Yang , Mårten S G Ahlquist , Licheng Sun †,⊥,#,*
PMCID: PMC10236490  PMID: 37196315

Abstract

graphic file with name ja3c03415_0011.jpg

Many enzymes use adaptive frameworks to preorganize substrates, accommodate various structural and electronic demands of intermediates, and accelerate related catalysis. Inspired by biological systems, a Ru-based molecular water oxidation catalyst containing a configurationally labile ligand [2,2′:6′,2″-terpyridine]-6,6″-disulfonate was designed to mimic enzymatic framework, in which the sulfonate coordination is highly flexible and functions as both an electron donor to stabilize high-valent Ru and a proton acceptor to accelerate water dissociation, thus boosting the catalytic water oxidation performance thermodynamically and kinetically. The combination of single-crystal X-ray analysis, various temperature NMR, electrochemical techniques, and DFT calculations was utilized to investigate the fundamental role of the self-adaptive ligand, demonstrating that the on-demand configurational changes give rise to fast catalytic kinetics with a turnover frequency (TOF) over 2000 s–1, which is compared to oxygen-evolving complex (OEC) in natural photosynthesis.

Introduction

The design of catalysts that rival the proficiency of metalloenzymes is an intense research topic of coordination chemistry.1,2 Enzymes catalyze reactions in a dynamic manner, binding substrates at the labile coordination site and releasing products by interconverting the conformations on different time scales (Figure 1a).35 Several crystallographic and time-resolved spectroscopic techniques have disclosed dynamic ligand exchanges in the vicinity of the catalytic site during photosynthesis,6,7 nitrogen fixation,8 oxygen reduction,9,10 DNA synthesis/cleavage,1113 etc.14 For example, in the Kok cycle of photosystem II (PSII), the substrate water would rearrange from Ca to pentacoordinate Mn in the S2 → S3 transition accompanied by the oxidation of Mn from III to IV,6,1520 during which the Ca-ligated D1-Glu189 residue moves away accordingly and makes space for the substrate water coordination (Figure 1b). The detached D1-Glu189 or/and other residues serve as proton acceptors, transferring protons from the catalytic center to bulk water.6,21,22

Figure 1.

Figure 1

(a) Rearrangement of first and second coordination spheres during enzymatic catalysis. (b) Transition from S2 to S3 during the oxygen-evolving Kok cycle (Ox stands for OH or O). (c) Proposed self-adaptive configurations during water oxidation catalysis by Ru-tds. Axial ligands are omitted for clarity.

To mimic the function of the adaptive architecture in PSII, conformationally flexible ligands with the ability to accommodate the structural and electronic demands of the different intermediates have been successfully applied in artificial water oxidation catalysts and have resulted in a few elegant examples of seven-coordination phenomena in Ru-bda and Ru-tda (Scheme 1, bda = 2,2′-bipyridine-6,6′-dicarboxylate, tda = [2,2′:6′,2″-terpyridine]-6,6″-dicarboxylate) catalysts that display enhanced activity comparable to the Mn4CaOx cluster of PSII.23,24 The classic Ru-bda system mediates O–O bond formation via the interaction of two metal-oxyl species (I2M), which is highly dependent on the interaction between the catalysts.25 Modifications on distal ligands to preorganize the substrate water can promote an alternative water nucleophilic attack (WNA) pathway, where the preorganized water network serves as a base to facilitate the proton transfer process.2628 Another intriguing strategy is to introduce intramolecular proton acceptors by rearrangements of the coordination conformations, which is skilfully illustrated by Ru-tda and Ru-tpa type water oxidation catalysts (Scheme 1, tpa = 2,2′:6′,2″-terpyridine-6,6″-diphosphonate).2931 However, installations of proton acceptor at the second coordination sphere inevitably lead to a competitive coordination with substrate water, making it impossible to fully leverage the catalytic site.29,32 Therefore, the coordination ability of the proton relay unit needs to be negotiated with the water molecule to lower the energy required for substrate binding and activation.33,34

Scheme 1. Structures of Water Oxidation Catalysts Discussed in the Paper and Ru-tds.

Scheme 1

Organosulfonates are a relatively unexplored type of ligand among the reported water oxidation catalysts3540 due to their relatively weak coordination capabilities,4143 while they offer potential applications in adaptive chemistry. As a proton acceptor, sulfonate is also able to accept up to six hydrogen bonds with the lone pairs of oxygen. Indeed, replacing the dicarboxylates of Ru-bda with disulfonates enables a 40-fold increase in water oxidation efficiency (Scheme 1, Ru-bds, bds = 2,2′-bipyridine-6,6′-disulfonate).44 Introduction of a remote sulfonate at the second coordination sphere also leads to a boosted performance in the context of higher-onset-potential Ru-tpy-type catalysts (around 800 mV, tpy = terpyridine), allowing for excellent performance under both acidic and basic conditions (Scheme 1, Grotjahn-Cat).45,46 Ideally, simultaneously introducing sulfonates to both the first and second coordination spheres can enrich the electron density of the metal, promote substrate binding, and accelerate the proton transfer process (Figure 1c), enabling fast water oxidation kinetics under a mild driving force.

In this work, a bioinspired catalyst with an adaptive architecture, Ru-tds (tds = [2,2′:6′,2″-terpyridine]-6,6″-disulfonate, Scheme 1), is designed to resemble enzymes, where the ligand can satisfy the varied electronic and geometric requirements of catalytic intermediates through dynamic sulfonate coordination/de-coordination. The catalyst achieves high TOFs over 2000 s–1 with a mild onset potential of 530 mV and an overpotential of 620 mV under neutral conditions, which is compared to the Mn4CaOx cluster of PSII. Spectroscopic and kinetic studies in concert with computational results reveal that the proton transfer events in the catalytic cycle are fast enough; as such, the rate-determining step (RDS) shifts to the substrate binding process via aqua-sulfonate ligand exchange.

Experimental Section

Synthesis and Characterization

The ligand [2,2′:6′,2″-terpyridine]-6,6″-disulfonic acid (H2tds) was synthesized in two steps as described in the Supporting Information. In short, 6,6″-dibromo-2,2′:6′,2″-terpyridine was initially transformed into [2,2′:6′,2″-terpyridine]-6,6″-dithiol via a nucleophilic aromatic substitution reaction, followed by oxidation of the dithiol to disulfonic acid. Complex Ru-tds was prepared via a one-pot reaction, i.e., refluxing H2tds, [Ru(DMSO)4Cl2], and pyridine in ethanol under N2. The desired catalyst was isolated via column chromatography and characterized by nuclear magnetic resonance (NMR) and high-resolution mass spectrometry (HRMS) (Figures S1–S6). Two byproducts were also isolated by column chromatography and characterized by 1H NMR, which are tentatively assigned to Ru(tds)2 and Ru(tds)(py)(DMSO), respectively (Figure S22).

Results and Discussion

Two conformations of the single-crystal structure, i.e., RuII(tds-κ-N3O)Py2 and RuII(tds-κ-N3O2)Py2, were obtained in different batches of crystal growth (Figure 2, CCDC: 2209276 and 2209277). Complex RuII(tds-κ-N3O)Py2 features a distorted octahedral geometry with a dangling sulfonate at the second coordination sphere. The Npyridine–Ru–Osulfonate angle is 124.3°, which is slightly larger than that of carboxyl-containing analogues Ru-tda and Ru-bda(29,47) and much larger than that of phosphonate-containing (Ru-tpa) and sulfonate-containing (Ru-bds) analogues (Figure 2a and Table 1).30,44 The large angle can serve as the site for water binding, activation, and O–O bond formation. Complex RuII(tds-κ-N3O2)Py2 exhibits a pentagonal bipyramidal coordination geometry with splitting positions of oxygen atoms (O1a/1b and O3a/3b, Figure 2b), suggesting an alternate position of the sulfonate. To the best of our knowledge, this is the first isolated pseudo-seven-coordinated Ru(II) complex. It should be noted that the seven-coordinate Ru(II) complex (20-electron rule) is thermodynamically unstable due to the violation of the 18-electron rule. Collectively considering the long Ru–O distance (2.30–2.35 Å) and the splitting positions, it is hypothesized that sulfonates in both RuII(tds-κ-N3O)Py2 and RuII(tds-κ-N3O2)Py2 are weakly bound. The weak sulfonate coordination can promote substrate binding in the subsequent catalysis step via ligand exchange.

Figure 2.

Figure 2

Single-crystal structures of (a) RuII(tds-κ-N3O)Py2 and (b) RuII(tds-κ-N3O2)Py2 with thermal ellipsoids at 50% probability. Hydrogen atoms and solvent molecules are omitted for clarity. (c) VT 1H NMR spectra of Ru-tds in CD3OD/D2O (v/v = 4/1).

Table 1. Crystal and Catalysis Data for Ru-tds and Other State-of-the-art Reference Catalysts.

graphic file with name ja3c03415_0010.jpg

  Ru-tds Grotjahn-Cat45 Ru-tpa(30) Ru-tda(29) Ru-bda(47) Ru-bds(44)
Ru–O (Å)a 2.35 2.18 2.19 2.14–2.20 2.17–2.21 2.19–2.21
O–Ru–O(N) (°)a 124.3   114.6 122.4 123.0 114.7–115.1
onset potential (V)b 1.76 2.02 1.76 1.70–1.80 1.55 1.65
overpotential (V)b 0.62 0.80 0.69 0.64 0.21/0.67 0.62
TOF (s–1) method Ac 2239 ± 311          
method Bd 3242 2598        
method Ce 4195       300 12 900
TOFmax (s–1)f 12 000   16 000 8000    
mechanismg WNA WNA WNA WNA I2M I2M
a

Data for bond length and angle extracted from their crystal structures.

b

Values of onset potential vs RHE extracted from their CVs at pH 7; values of overpotential (Ecat/2) extracted from their CVs at pH 7 according to the suggested method in the literature,49 and the two values for Ru-bda are extracted from two “catalytic plateau,” respectively.44,50

c

TOF stands for turnover frequency, values extracted from their CVs at pH 7 according to eq 3 in the Supporting Information.

d

Values estimated according to eq 2 in the Supporting Information and the reference (45) at a scan rate of 0.01 V s–1 at pH 7.

e

Values estimated according to eq 4 in the Supporting Information and the reference (44) at a scan rate of 0.1 V s–1 at pH 7.

f

TOFmax stands for turnover frequency maximum that was estimated by the foot of the wave analysis29 at pH 7 (for details, see Figure S14). For Ru-tpa and Ru-tda, the values were calculated after catalyst activation.

g

WNA stands for water nucleophilic attack, and I2M stands for interaction of two metal-oxo.

The 1H NMR spectra show that Ru-tds maintained its symmetry in solution, in contrast to the asymmetrical conformation found in the single crystal. The 1H NMR spectrum shows only three signals for the axial ligands and five signals for the equatorial ligand at room temperature (298 K, Figure 2c), which suggests a fast dynamic coordination behavior of the sulfonate groups. The dynamic coordination was then investigated by recording 1H NMR spectra at lower temperatures in a mixed solvent (CD3OD/D2O, v/v = 4/1) until its freezing point was reached (ca. 198 K). In contrast to the observations for the Ru-tpa analogue containing two phosphate groups,30 the spectra did not change with temperature (Figure 2c). Even below 200 K, one set of signals can be observed for the ligands, indicating a fast chemical exchange on the NMR chemical shift time scale, that is, a fast coordination/de-coordination of the sulfonate groups. This rate is much faster than that observed for phosphate coordination/de-coordination in the Ru-tpa system, where separate signals of the ligand can be observed below 253 K. In our view, the line broadening of the signals observed below 218 K is due to the fast transverse relaxation (T2) caused by the increased viscosity of the solvent.

Electrochemical Studies

The electrochemical properties of Ru-tds were investigated by cyclic voltammetry (CV) at pH 7.0 in a 0.1 M phosphate buffer solution containing 1% CF3CH2OH. A pH-independent oxidation peak (Figure 3c) was observed at Eox = 1.1 V, which can be assigned to a pure electron transfer process and will be discussed later. The CVs at different scan rates indicate that Eox is a diffusion-controlled electrochemical process, demonstrating a linear relationship between peak currents and the square root of scan rates (Figures S10 and S11). Subsequently, a substantial enhancement of the catalytic peak at 1.35 V vs NHE (onset potential) was observed, followed by a large catalytic current density of 4.85 mA cm–2 at 1.7 V by using a boron-doped diamond electrode (BDD, 0.0314 cm2, Figure 3a) as the working electrode. Under the same experimental conditions, the current density of the reference catalyst Ru-bda is 4.5 times lower (1.06 mA cm–2, Figure 3a) than that of Ru-tds.

Figure 3.

Figure 3

(a) CVs without background subtraction of 0.13 mM Ru-tds and Ru-bda at pH 7.0 in a 0.1 M phosphate buffer solution containing 1% CF3CH2OH, scan rate = 100 mV s–1, working electrode: BDD. (b) Potential vs pH diagram for Ru-tds in aqueous buffer solutions containing 1% CF3CH2OH, in which the potentials of RuV/IV were measured at a current of 3.7 μA from their LSVs. (c) CVs without background subtraction of Ru-tds at different potential windows, working electrode: BDD. (d) Negative-scan CV without background subtraction from 1.3 V, working electrode: BDD.

The differential pulse voltammograms (DPVs) of Ru-tds at various pH values were measured to study the proton and electron transfer processes during water activation and O–O bond formation (Figures 3b and S7). A pH-independent oxidation process at 1.10 V appears in the pH range of 5–10. Since Ru species generated during this process cannot trigger water oxidation, we further extracted the potentials of higher valent species from their catalytic peaks (details for the potential determination can be found in the Note below Figure S7). The catalytic currents are pH-dependent with a slope of −0.063 V/pH (Figure 3b, blue) in the pH range of 5–7, suggesting that a 1H+/1e transfer process occurs before O–O bond formation, most likely on RuV/IV. Accordingly, the first procedure at 1.1 V should be a 2e removal process as the catalyst precursor is in the RuII state. Our conclusion is in agreement with the fact that the analogue catalyst Ru-tda precursor (without an aqua ligand) can be oxidized to the RuIV state around 1.1 V in the same pH window.29

The water activation mechanism by Ru-tds is proposed in Scheme 2a based on the electrochemical data mentioned above. RuII-tds is initially oxidized to RuIV-tds, followed by the formation of seven-coordinated RuIV(OH)-tds as suggested by DFT calculation that will be discussed later (Figures 5, S20, and S21). Sulfonate is generally regarded as a poor ligand in terms of coordination ability, as indicated by its role in most metal aqua complexes being noncoordinating counter ions.41 Consequently, the Ru-aqua complex can be readily formed via ligand exchange. RuIV(OH)-tds then undergoes a 1H+/1e transfer process to generate RuV(O)-tds species.

Scheme 2. Proposed Water Oxidation Pathways of (a) Adaptive Catalysis by Ru-tds and (b) Semiadaptive Catalysis by the Analogue Catalyst,29.

Scheme 2

Axial ligands are omitted for clarity.

Figure 5.

Figure 5

Energy profiles of ligand exchange on RuIV at pH 7.0 using H2PO4 as the base. The units of energies are kcal mol–1.

Two reduction peaks appear at Ere1 = 1.02 V and Ere2 = 0.67 V, respectively, in the reverse CV scan (Figure 3c), which are assigned to consecutive one-electron reduction of the Ru-aqua species generated at higher potentials. This assumption is supported by the following evidence. First, reduction peaks only manifest once the RuIV state is reached (Figure 3c, purple), demonstrating that the reduction waves Ere1 and Ere2 are connected to the oxidation wave Eox. Second, the peak currents at Ere1 and Ere2 decrease as the scan window narrows (Figure 3c, red and blue), while the peak current at Eox remains the same, suggesting that the Ru-aqua complex is preferred at higher potentials and longer time scales. Third, the backward scan in Figure 3d shows that the two reduction peaks gradually vanish in the subsequent CVs in a smaller window, indicating that these two signals originate from diffusible active species rather than species adsorbed on the electrode surface due to catalyst decomposition. Fourth, the in situ formation of new active Ru-aqua species is indicated by the curve crossing between the forward and backward scans29,48 under a high scan rate (100 mV s–1, Figure 3a). This crossover disappears as the scan rate decreases (10 mV s–1, Figure S15) because the longer time scale results in the maximum conversion of Ru-tds to the corresponding catalytically active Ru-aqua species in the electrical double layer. Fifth, the potentials of Ere2 are pH-dependent with a slope of −0.051 V/pH (Figure S16) in the pH range of 5–8, validating the formation of Ru-aqua species. Finally, the analogue catalyst Ru(H2O)-tda also showed reduction potentials similar to those of RuIII/IV and RuII/III (0.93 and 0.72 V) at pH 7.29

Catalytic Performance

The generation of oxygen was confirmed by controlled potential electrolysis (1.7 V vs NHE) at pH 7 and monitored by a pressure transducer. The average Faradaic efficiency for water oxidation is over 92%, indicating that the majority of gas produced is oxygen (Figure S19). The linear relationship between catalytic current and catalyst concentration (Figures S8 and S9) and scan rate-independent catalytic current (Figures S12 and S13) enables us to evaluate the catalytic TOF by eq 3 in the Supporting Information, which provides a reliable method to compare the TOF of Ru-tds with the majority of reported catalysts.5154 The TOF value of 2239 ± 311 s–1 is compared to that of OEC in PSII and among the highest activities reported thus far for Ru- and non-noble metal-based catalysts.55 In addition, other modified methods are also used to fairly compare activities with reported state-of-the-art catalysts. (1) The sulfonate at the first coordination sphere is designed to enrich the electron density of Ru and to obtain a lower onset potential and overpotential. Indeed, the lower onset potential/overpotential and higher TOF are attained for Ru-tds compared to Grotjahn-Cat (Tables 1 and S2), in which the sulfonate is located only at the second coordination sphere.45 (2) We also estimated the TOFmax value of 12 000 s–1 according to the foot of the wave analysis (FOWA, Figure S14), however, this method assumes a scenario in which no side phenomena are operative. Instead, Ru-tds attained a comparable current density at a lower concentration than Ru-tda (0.13 vs 0.45 mM29), demonstrating that sulfonate at the second coordination sphere has a superior capacity for proton transfer than carboxylate. More importantly, the aqua-carboxylate exchange kinetics are slow (step 2″, Scheme 2b); as a consequence, an electrolysis-based activation procedure to generate the catalytically active species is necessary for Ru-tda (step 2′, Scheme 2b).29 This procedure is not required for Ru-tds due to the flexible coordination ability of sulfonates (step 2, Scheme 2a). The negatively charged sulfonates can re-coordinate to stabilize the charged metal center after oxygen release for Ru-tds (step 6, Scheme 2a) to close the adaptive catalytic cycle, whereas the semiadaptive Ru-tda catalyst either generated new active species with only neutral ligands (step 7, Scheme 2b) or returned to the catalyst precursor (step 6′, Scheme 2b). After long-term electrolysis, a new oxidation signal at 0.9 V with a relatively weak peak current appeared (inset, Figure 6a), suggesting a possible alternative pathway to generate RuII(OH2)-tds (Scheme S1). (3) The Ru-tpa also shared a semiadaptive catalytic process and underwent more complicated structural transformations to generate the catalytically active species Ru-tpaO and Ru-bpc and other catalytically inactive species under high potentials (Table S2).30,31,56 (4) Although Ru-bds exhibits higher catalytic activity than Ru-tds, this is because Ru-bds catalyzes water oxidation via the I2M mechanism, where the TOF is proportional to the catalyst concentration (TOFI2M = k [cat], TOFWNA = k).23 Catalyst with the WNA mechanism is more promising for practical applications, such as immobilization on the electrode surface.

Figure 6.

Figure 6

(a) 1st and 200th CV scans without background subtraction of Ru-tds at pH 7.0 in a 0.1 M phosphate buffer solution containing 1% CF3CH2OH, scan rate = 100 mV s–1, [Ru-tds] = 0.13 mM, working electrode: BDD; inset: enlargement of the 0.2–1.4 V range. (b) Controlled potential electrolysis (1.7 V) in a 0.1 M phosphate buffer solution containing 1% CF3CH2OH with Ru-tds (0.13 mM, red) and only with the electrode (gray), working electrode: GC; inset: bubble formation on the electrode surface.

Cation, Anion, and Kinetic Isotope Effects

More insights into the catalytic mechanism of Ru-tds were obtained by cation, anion, and kinetic isotope effect (KIE) analysis. The catalytic current for water oxidation is linearly dependent on the concentration of the catalyst (Figures S8 and S9), suggesting that the RDS should take place on a single site in accordance with a WNA pathway. Since the O–H bond cleavage was implicated in steps 3 and 5, as proposed in Scheme 2a, the possibility that these steps were the RDS can be excluded according to a secondary KIE generated from the CV scans (Figure 4a). The negligible pH-dependent catalytic currents (icat/ip) further support that the proton transfer is not involved in the RDS (Figure S17). Additionally, due to the varying solvation strengths (Li+ > Na+ > K+, Figure 4b), the nucleophilic attack ability of water is correlated with the types of cations utilized in electrolytes.57,58 The CVs were then measured in 0.1 M LiPi, NaPi, and KPi buffer solutions, as shown in Figure 4d, with insignificant cation effects in the potential range of 1.95–2.05 V against RHE, indicating that step 4 is not involved in RDS. The oxygen release (step 6) is significantly less demanding than the O–O bond formation from a free energy point of view.59 Collectively, the RDS should be the formation of Ru-aqua species (step 2, Scheme 2a). It is interesting to note that we also discovered relatively high KIE values in a backward scan LSV (Figure S18) and relatively obvious cation effects at the potentials of 2.10 and 2.20 V (Figure 4d), suggesting that the ligand exchange to produce Ru-aqua species is accelerated under the higher potential and longer time scale, resulting in the RDS being transferred somehow from step 2 to the step where the O–O bond formation takes place. Besides, as the buffer concentration increased from 0.01 to 0.10 M, base-enhanced water oxidation is observed (Figure 4c). The catalytic current reached a plateau as the buffer concentration increased further, whereas the anion can affect the catalytic activities of Ru-bda and Ru-bds over a broader range (0.01–0.20 M) due to the different RDS involved (proton-coupled oxidation step under the same conditions).44,50 Therefore, the involvement of the buffer in step 2 contributes to the faster aqua-sulfonate ligand exchange. In summary, the RDS can be accelerated via the maximum formation of Ru-aqua species in the presence of concentrated buffer solution under higher potentials and longer time scales.

Figure 4.

Figure 4

(a) CVs without background subtraction of 0.13 mM Ru-tds in a 0.1 M phosphate buffer solution in H2O and D2O, (pH = 7 and pD = 7.87) containing 1% CF3CH2OH, scan rate = 20 mV s–1, working electrode: BDD. (b) Schematic diagram of potential anion and cation effects for water oxidation. (c) CVs without background subtraction of 0.13 mM Ru-tds in phosphate buffer solution with various concentrations (pH 7.1) containing 1% CF3CH2OH, scan rate = 20 mV s–1, working electrode: BDD; inset: enlargement of the 1.0–1.3 V range (upper) and a plot of icat/ip vs [phosphate] (bottom). Ionic strength kept at 0.5 M with NaClO4. (d) CVs without background subtraction of 0.13 mM Ru-tds in 0.1 M KPi, NaPi, and LiPi, working electrode: BDD; inset: comparison of icat/ip at different potentials, scan rate = 20 mV s–1.

DFT Study on RDS

DFT calculations were performed to elucidate the high performance of the electrochemical-driven water oxidation by Ru-tds at pH 7.0 (Figure 5). The calculated potential from the optimized 6-coordinate RuII species to 7-coordinate RuIV (Figure S20) is 1.25 eV, which is in good agreement with the experimental redox potential. In the 7-coordinate RuIV, the O–Ru–O angle is only 68.6°, which could increase the barrier of the aqua ligand coordination to the Ru atom. This step is also tested as the RDS from the experiment; therefore, the detailed calculations were focused on this step. At an aqueous solution, the water coordination is endergonic with an activation free energy of 19.3 kcal mol–1 (Figure S21). While in the phosphate buffer, H2PO4 (dominant species in pH 7.0 phosphate buffer) could stabilize the transition state by forming the H-bond and further taking the proton from a water molecule, leading to a much lower activation free energy of 13.9 kcal mol–1. The reaction is exergonic and generates a structure, where the OH forms the H-bond with the sulfonate group. This low activation free energy is consistent with a high TOF value. Hence, the buffer-promoted performance of Ru-tds can be ascribed to the fact that the buffer molecules facilitate the deprotonation of water to form OH, accelerating the kinetics of the ligand exchange.

Catalyst Stability

The stability of Ru-tds for water oxidation is monitored via repetitive CV and controlled potential electrolysis (CPE) under neutral conditions (Figure 6). To test steady-state durability, the catalyst was first subjected to 5-cycle CV scans to generate enough Ru-aqua species. Figure 6a shows that the catalytic current increases slightly after 200 cycles of CV scan, which is due to the increased amount of Ru-aqua active species in the electrical double layer, as indicated by the decreased intensity of Eox and increased intensity of Ere1 and Ere2. Following numerous CV scans, the electrode was removed from the solution, rinsed with water, and inserted into a brand-new electrolyte solution devoid of catalyst. The disappearance of redox and catalytic signals (gray, Figure 6a) indicates that no active species were deposited on the electrode surface throughout the whole stability tests. The CPE at 1.7 V vs NHE demonstrates that the current decreases gradually over time, which is a result of oxygen bubble formation on the electrode surface (Figure 6b). The catalytic current can be recovered by gently tickling the electrode to dislodge the bubbles. Taken together, these findings strongly suggest that Ru-tds functions as a reliable and efficient homogeneous molecular water oxidation catalyst at neutral pH.

Conclusions

To conclude, the development of water oxidation catalysts that can mimic the dynamic catalytic nature of enzymes presents a timely challenge. The introduction of sulfonates to the ruthenium complex created labile coordination spheres: (1) the coordination of negatively charged sulfonates enriches the electron density of ruthenium, thus thermodynamically stabilizing the positively charged catalytic site at the initial state; (2) the dynamic sulfonate coordination/de-coordination creates an open site for water binding and enables the immediate formation of the Ru-aqua active species via aqua-sulfonate ligand exchange without an extra driving force, which is indispensable for the subsequent O–O bond formation; (3) substrate water absorption also results in a dangling and non-coordinated sulfonate for the kinetic acceleration of the proton transfer process; and (4) the dangling sulfonate re-coordinates to the ruthenium after the product oxygen releases to further stabilize the charged catalytic site. Consequently, high TOFs (2000–4000 s–1) were obtained with a mild onset potential of 530 mV and an overpotential of 620 mV. The dynamic nature of the Ru-tds catalyst has been proven by the combination of single-crystal X-ray analysis, VT NMR, electrochemical techniques, and theoretical studies. The introduction of labile sulfonate may provide a general strategy for homogeneous and even heterogeneous water oxidation catalysis and other related proton-coupled electron transfer reactions.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (22088102), the Swedish Research Council (no. 2017-00935), and the Starting-up Package of Westlake University. S.Z. thanks the financial support from the Swedish Research Council (the Vetenskapsrådet international postdoc grant). N.S. acknowledges the financial support from the National Natural Science Foundation of China (22006108). All calculations were performed on resources provided by the Swedish National Infrastructure for Computing (SNIC) (allocations SNIC 2021/5-591 and SNIC 2021/6-345) at the National Supercomputing Center in Linköping, Sweden. The authors thank Dr. Yuye Zhou (KTH) and Dr. Yinjuan Chen (Instrumentation and Service Centre for Molecular Sciences at Westlake University) for supporting in MS measurement and Dr. Yu Guo (Westlake University) for discussion.

Data Availability Statement

Crystallographic data for the structures reported in this Article have been deposited at the Cambridge Crystallographic Data Centre under deposition numbers CCDC 2209276 and 2209277. Copies of the data can be obtained free of charge from www.ccdc.cam.ac.uk/structures/. Other data that support the findings of this study are available from the corresponding author upon reasonable request.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c03415.

  • Experimental procedures; compounds synthesis and characterization (NMR, HRMS, and crystal data); electrochemical data; and DFT calculations (PDF)

Author Contributions

T.L., S.Z., and N.S. contributed equally to this work.

The authors declare no competing financial interest.

Supplementary Material

References

  1. Wiester M. J.; Ulmann P. A.; Mirkin C. A. Enzyme Mimics Based Upon Supramolecular Coordination Chemistry. Angew. Chem., Int. Ed. 2011, 50, 114–137. 10.1002/anie.201000380. [DOI] [PubMed] [Google Scholar]
  2. Raynal M.; Ballester P.; Vidal-Ferran A.; van Leeuwen P. W. N. M. Supramolecular catalysis. Part 2: artificial enzyme mimics. Chem. Soc. Rev. 2014, 43, 1734–1787. 10.1039/C3CS60037H. [DOI] [PubMed] [Google Scholar]
  3. Callender R.; Dyer R. B. The Dynamical Nature of Enzymatic Catalysis. Acc. Chem. Res. 2015, 48, 407–413. 10.1021/ar5002928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Ayipo Y. O.; Osunniran W. A.; Babamale H. F.; Ayinde M. O.; Mordi M. N. Metalloenzyme mimicry and modulation strategies to conquer antimicrobial resistance: Metal-ligand coordination perspectives. Coord. Chem. Rev. 2022, 453, 214317 10.1016/j.ccr.2021.214317. [DOI] [Google Scholar]
  5. Zhao M.; Wang H.-B.; Ji L.-N.; Mao Z.-W. Insights into metalloenzyme microenvironments: biomimetic metal complexes with a functional second coordination sphere. Chem. Soc. Rev. 2013, 42, 8360–8375. 10.1039/c3cs60162e. [DOI] [PubMed] [Google Scholar]
  6. Kern J.; Chatterjee R.; Young I. D.; Fuller F. D.; Lassalle L.; Ibrahim M.; Gul S.; Fransson T.; Brewster A. S.; Alonso-Mori R.; Hussein R.; Zhang M.; Douthit L.; de Lichtenberg C.; Cheah M. H.; Shevela D.; Wersig J.; Seuffert I.; Sokaras D.; Pastor E.; Weninger C.; Kroll T.; Sierra R. G.; Aller P.; Butryn A.; Orville A. M.; Liang M.; Batyuk A.; Koglin J. E.; Carbajo S.; Boutet S.; Moriarty N. W.; Holton J. M.; Dobbek H.; Adams P. D.; Bergmann U.; Sauter N. K.; Zouni A.; Messinger J.; Yano J.; Yachandra V. K. Structures of the intermediates of Kok’s photosynthetic water oxidation clock. Nature 2018, 563, 421–425. 10.1038/s41586-018-0681-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bao H.; Burnap R. L. Structural rearrangements preceding dioxygen formation by the water oxidation complex of photosystem II. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, E6139–E6147. 10.1073/pnas.1512008112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Chan M. K.; Kim J.; Rees D. C. The Nitrogenase FeMo-Cofactor and P-Cluster pair: 2.2 Å Resolution Structures. Science 1993, 260, 792–794. 10.1126/science.8484118. [DOI] [PubMed] [Google Scholar]
  9. Bhunia S.; Ghatak A.; Dey A. Second Sphere Effects on Oxygen Reduction and Peroxide Activation by Mononuclear Iron Porphyrins and Related Systems. Chem. Rev. 2022, 122, 12370–12426. 10.1021/acs.chemrev.1c01021. [DOI] [PubMed] [Google Scholar]
  10. Dubey K. D.; Shaik S. Cytochrome P450—The Wonderful Nanomachine Revealed through Dynamic Simulations of the Catalytic Cycle. Acc. Chem. Res. 2019, 52, 389–399. 10.1021/acs.accounts.8b00467. [DOI] [PubMed] [Google Scholar]
  11. Hwang W.; Yoo J.; Lee Y.; Park S.; Hoang P. L.; Cho H.; Yu J.; Hoa Vo T. M.; Shin M.; Jin M. S.; Park D.; Hyeon C.; Lee G. Dynamic coordination of two-metal-ions orchestrates λ-exonuclease catalysis. Nat. Commun. 2018, 9, 4404 10.1038/s41467-018-06750-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Gao Y.; Yang W. Capture of a third Mg2+ is essential for catalyzing DNA synthesis. Science 2016, 352, 1334–1337. 10.1126/science.aad9633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Shi Y.; Hellinga H. W.; Beese L. S. Interplay of catalysis, fidelity, threading, and processivity in the exo- and endonucleolytic reactions of human exonuclease I. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 6010–6015. 10.1073/pnas.1704845114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Holm R. H.; Kennepohl P.; Solomon E. I. Structural and Functional Aspects of Metal Sites in Biology. Chem. Rev. 1996, 96, 2239–2314. 10.1021/cr9500390. [DOI] [PubMed] [Google Scholar]
  15. Siegbahn P. E. M. Structures and Energetics for O2 Formation in Photosystem II. Acc. Chem. Res. 2009, 42, 1871–1880. 10.1021/ar900117k. [DOI] [PubMed] [Google Scholar]
  16. Cox N.; Retegan M.; Neese F.; Pantazis D. A.; Boussac A.; Lubitz W. Electronic structure of the oxygen-evolving complex in photosystem II prior to O-O bond formation. Science 2014, 345, 804–808. 10.1126/science.1254910. [DOI] [PubMed] [Google Scholar]
  17. Dau H.; Haumann M. Considerations on the mechanism of photosynthetic water oxidation – dual role of oxo-bridges between Mn ions in (i) redox-potential maintenance and (ii) proton abstraction from substrate water. Photosynth. Respir. 2005, 84, 325–331. 10.1007/s11120-004-7080-2. [DOI] [PubMed] [Google Scholar]
  18. Ibrahim M.; Fransson T.; Chatterjee R.; Cheah M. H.; Hussein R.; Lassalle L.; Sutherlin K. D.; Young I. D.; Fuller F. D.; Gul S.; Kim I.-S.; Simon P. S.; de Lichtenberg C.; Chernev P.; Bogacz I.; Pham C. C.; Orville A. M.; Saichek N.; Northen T.; Batyuk A.; Carbajo S.; Alonso-Mori R.; Tono K.; Owada S.; Bhowmick A.; Bolotovsky R.; Mendez D.; Moriarty N. W.; Holton J. M.; Dobbek H.; Brewster A. S.; Adams P. D.; Sauter N. K.; Bergmann U.; Zouni A.; Messinger J.; Kern J.; Yachandra V. K.; Yano J. Untangling the sequence of events during the S2→S3 transition in photosystem II and implications for the water oxidation mechanism. Proc. Natl. Acad. Sci. U.S.A. 2020, 117, 12624–12635. 10.1073/pnas.2000529117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Kim C. J.; Debus R. J. Evidence from FTIR Difference Spectroscopy That a Substrate H2O Molecule for O2 Formation in Photosystem II Is Provided by the Ca Ion of the Catalytic Mn4CaO5 Cluster. Biochemistry 2017, 56, 2558–2570. 10.1021/acs.biochem.6b01278. [DOI] [PubMed] [Google Scholar]
  20. Kim C. J.; Debus R. J. One of the Substrate Waters for O2 Formation in Photosystem II Is Provided by the Water-Splitting Mn4CaO5 Cluster’s Ca2+ Ion. Biochemistry 2019, 58, 3185–3192. 10.1021/acs.biochem.9b00418. [DOI] [PubMed] [Google Scholar]
  21. Suga M.; Akita F.; Sugahara M.; Kubo M.; Nakajima Y.; Nakane T.; Yamashita K.; Umena Y.; Nakabayashi M.; Yamane T.; Nakano T.; Suzuki M.; Masuda T.; Inoue S.; Kimura T.; Nomura T.; Yonekura S.; Yu L.-J.; Sakamoto T.; Motomura T.; Chen J.-H.; Kato Y.; Noguchi T.; Tono K.; Joti Y.; Kameshima T.; Hatsui T.; Nango E.; Tanaka R.; Naitow H.; Matsuura Y.; Yamashita A.; Yamamoto M.; Nureki O.; Yabashi M.; Ishikawa T.; Iwata S.; Shen J.-R. Light-induced structural changes and the site of O=O bond formation in PSII caught by XFEL. Nature 2017, 543, 131–135. 10.1038/nature21400. [DOI] [PubMed] [Google Scholar]
  22. Askerka M.; Wang J.; Vinyard D. J.; Brudvig G. W.; Batista V. S. S3 State of the O2-Evolving Complex of Photosystem II: Insights from QM/MM, EXAFS, and Femtosecond X-ray Diffraction. Biochemistry 2016, 55, 981–984. 10.1021/acs.biochem.6b00041. [DOI] [PubMed] [Google Scholar]
  23. Zhang B.; Sun L. Ru-bda: Unique molecular water-oxidation catalysts with distortion induced open site and negatively charged ligands. J. Am. Chem. Soc. 2019, 141, 5565–5580. 10.1021/jacs.8b12862. [DOI] [PubMed] [Google Scholar]
  24. Matheu R.; Ertem M. Z.; Gimbert-Suriñach C.; Sala X.; Llobet A. Seven coordinated molecular ruthenium–water oxidation catalysts: a coordination chemistry journey: focus review. Chem. Rev. 2019, 119, 3453–3471. 10.1021/acs.chemrev.8b00537. [DOI] [PubMed] [Google Scholar]
  25. Duan L.; Wang L.; Li F.; Li F.; Sun L. Highly Efficient Bioinspired Molecular Ru Water Oxidation Catalysts with Negatively Charged Backbone Ligands. Acc. Chem. Res. 2015, 48, 2084–2096. 10.1021/acs.accounts.5b00149. [DOI] [PubMed] [Google Scholar]
  26. Liu T.; Li G.; Shen N.; Wang L.; Timmer B. J. J.; Zhou S.; Zhang B.; Kravchenko A.; Xu B.; Ahlquist M. S. G.; Sun L. Isolation and Identification of Pseudo Seven-Coordinate Ru(III) Intermediate Completing the Catalytic Cycle of Ru-bda Type of Water Oxidation Catalysts. CCS Chem. 2022, 4, 2481–2490. 10.31635/ccschem.021.202101159. [DOI] [Google Scholar]
  27. Liu T.; Li G.; Shen N.; Wang L.; Timmer B. J. J.; Kravchenko A.; Zhou S.; Gao Y.; Yang Y.; Yang H.; Xu B.; Zhang B.; Ahlquist M. S. G.; Sun L. Promoting Proton Transfer and Stabilizing Intermediates in Catalytic Water Oxidation via Hydrophobic Outer Sphere Interactions. Chem. - Eur. J. 2022, 28, e202104562 10.1002/chem.202104562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Noll N.; Krause A.-M.; Beuerle F.; Würthner F. Enzyme-like water preorganization in a synthetic molecular cleft for homogeneous water oxidation catalysis. Nat. Catal. 2022, 5, 867–877. 10.1038/s41929-022-00843-x. [DOI] [Google Scholar]
  29. Matheu R.; Ertem M. Z.; Benet-Buchholz J.; Coronado E.; Batista V. S.; Sala X.; Llobet A. Intramolecular Proton Transfer Boosts Water Oxidation Catalyzed by a Ru Complex. J. Am. Chem. Soc. 2015, 137, 10786–10795. 10.1021/jacs.5b06541. [DOI] [PubMed] [Google Scholar]
  30. Vereshchuk N.; Matheu R.; Benet-Buchholz J.; Pipelier M.; Lebreton J.; Dubreuil D.; Tessier A.; Gimbert-Suriñach C.; Ertem M. Z.; Llobet A. Second Coordination Sphere Effects in an Evolved Ru Complex Based on Highly Adaptable Ligand Results in Rapid Water Oxidation Catalysis. J. Am. Chem. Soc. 2020, 142, 5068–5077. 10.1021/jacs.9b11935. [DOI] [PubMed] [Google Scholar]
  31. Vereshchuk N.; Holub J.; Gil-Sepulcre M.; Benet-Buchholz J.; Llobet A. Fate of the Molecular Ru–Phosphonate Water Oxidation Catalyst under Turnover Conditions. ACS Catal. 2021, 11, 5240–5247. 10.1021/acscatal.0c05363. [DOI] [Google Scholar]
  32. Shatskiy A.; Bardin A. A.; Oschmann M.; Matheu R.; Benet-Buchholz J.; Eriksson L.; Kärkäs M. D.; Johnston E. V.; Gimbert-Suriñach C.; Llobet A.; Åkermark B. Electrochemically Driven Water Oxidation by a Highly Active Ruthenium-Based Catalyst. ChemSusChem 2019, 12, 2251–2262. 10.1002/cssc.201900097. [DOI] [PubMed] [Google Scholar]
  33. Liu T.; Sun L. Proton transfer regulating in catalytic water oxidation by Ru-complexes: second coordination sphere and beyond. Sci. Bull. 2023, 68, 854–856. 10.1016/j.scib.2023.04.012. [DOI] [PubMed] [Google Scholar]
  34. Liu T.; Sun L. The future challenges in molecular water oxidation catalysts. J. Energy Chem. 2022, 73, 643–645. 10.1016/j.jechem.2022.07.001. [DOI] [Google Scholar]
  35. Duan L.; Wang L.; Inge A. K.; Fischer A.; Zou X.; Sun L. Insights into Ru-based molecular water oxidation catalysts: Electronic and noncovalent-interaction effects on their catalytic activities. Inorg. Chem. 2013, 52, 7844–7852. 10.1021/ic302687d. [DOI] [PubMed] [Google Scholar]
  36. Neudeck S.; Maji S.; López I.; Meyer S.; Meyer F.; Llobet A. New Powerful and Oxidatively Rugged Dinuclear Ru Water Oxidation Catalyst: Control of Mechanistic Pathways by Tailored Ligand Design. J. Am. Chem. Soc. 2014, 136, 24–27. 10.1021/ja409974b. [DOI] [PubMed] [Google Scholar]
  37. Yang B.; Jiang X.; Guo Q.; Lei T.; Zhang L.-P.; Chen B.; Tung C.-H.; Wu L.-Z. Self-Assembled Amphiphilic Water Oxidation Catalysts: Control of O–O Bond Formation Pathways by Different Aggregation Patterns. Angew. Chem., Int. Ed. 2016, 55, 6229–6234. 10.1002/anie.201601653. [DOI] [PubMed] [Google Scholar]
  38. Yoshida M.; Kondo M.; Torii S.; Sakai K.; Masaoka S. Oxygen Evolution Catalyzed by a Mononuclear Ruthenium Complex Bearing Pendant SO3– Groups. Angew. Chem., Int. Ed. 2015, 54, 7981–7984. 10.1002/anie.201503365. [DOI] [PubMed] [Google Scholar]
  39. Yu F.; Poole D. III; Mathew S.; Yan N.; Hessels J.; Orth N.; Ivanović-Burmazović I.; Reek J. N. H. Control over Electrochemical Water Oxidation Catalysis by Preorganization of Molecular Ruthenium Catalysts in Self-Assembled Nanospheres. Angew. Chem., Int. Ed. 2018, 57, 11247–11251. 10.1002/anie.201805244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Koelewijn J. M.; Lutz M.; Dzik W. I.; Detz R. J.; Reek J. N. H. Reaction Progress Kinetic Analysis as a Tool To Reveal Ligand Effects in Ce(IV)-Driven IrCp*-Catalyzed Water Oxidation. ACS Catal. 2016, 6, 3418–3427. 10.1021/acscatal.6b00297. [DOI] [Google Scholar]
  41. Côté A. P.; Shimizu G. K. H. The supramolecular chemistry of the sulfonate group in extended solids. Coord. Chem. Rev. 2003, 245, 49–64. 10.1016/S0010-8545(03)00033-X. [DOI] [Google Scholar]
  42. Videnova-Adrabinska V. Coordination and supramolecular network entanglements of organodisulfonates. Coord. Chem. Rev. 2007, 251, 1987–2016. 10.1016/j.ccr.2007.03.018. [DOI] [Google Scholar]
  43. Shimizu G. K. H.; Vaidhyanathan R.; Taylor J. M. Phosphonate and sulfonate metal organic frameworks. Chem. Soc. Rev. 2009, 38, 1430–1449. 10.1039/b802423p. [DOI] [PubMed] [Google Scholar]
  44. Yang J.; Wang L.; Zhan S.; Zou H.; Chen H.; Ahlquist M. S. G.; Duan L.; Sun L. From Ru-bda to Ru-bds: a step forward to highly efficient molecular water oxidation electrocatalysts under acidic and neutral conditions. Nat. Commun. 2021, 12, 373 10.1038/s41467-020-20637-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Nash A. G.; Breyer C. J.; Vincenzini B. D.; Elliott G. I.; Niklas J.; Poluektov O. G.; Rheingold A. L.; Smith D. K.; Musaev D. G.; Grotjahn D. B. An Active-Site Sulfonate Group Creates a Fast Water Oxidation Electrocatalyst That Exhibits High Activity in Acid. Angew. Chem., Int. Ed. 2021, 60, 1540–1545. 10.1002/anie.202008896. [DOI] [PubMed] [Google Scholar]
  46. Kagalwala H. N.; Tong L.; Zong R.; Kohler L.; Ahlquist M. S. G.; Fan T.; Gagnon K. J.; Thummel R. P. Evidence for Oxidative Decay of a Ru-Bound Ligand during Catalyzed Water Oxidation. ACS Catal. 2017, 7, 2607–2615. 10.1021/acscatal.6b03278. [DOI] [Google Scholar]
  47. Duan L.; Fischer A.; Xu Y.; Sun L. Isolated Seven-Coordinate Ru(IV) Dimer Complex with [HOHOH] Bridging Ligand as an Intermediate for Catalytic Water Oxidation. J. Am. Chem. Soc. 2009, 131, 10397–10399. 10.1021/ja9034686. [DOI] [PubMed] [Google Scholar]
  48. Lee K. J.; McCarthy B. D.; Dempsey J. L. On decomposition, degradation, and voltammetric deviation: the electrochemist’s field guide to identifying precatalyst transformation. Chem. Soc. Rev. 2019, 48, 2927–2945. 10.1039/C8CS00851E. [DOI] [PubMed] [Google Scholar]
  49. Appel A. M.; Helm M. L. Determining the Overpotential for a Molecular Electrocatalyst. ACS Catal. 2014, 4, 630–633. 10.1021/cs401013v. [DOI] [Google Scholar]
  50. Song N.; Concepcion J. J.; Binstead R. A.; Rudd J. A.; Vannucci A. K.; Dares C. J.; Coggins M. K.; Meyer T. J. Base-enhanced catalytic water oxidation by a carboxylate–bipyridine Ru (II) complex. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 4935–4940. 10.1073/pnas.1500245112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Okamura M.; Kondo M.; Kuga R.; Kurashige Y.; Yanai T.; Hayami S.; Praneeth V. K.; Yoshida M.; Yoneda K.; Kawata S.; Masaoka S. A pentanuclear iron catalyst designed for water oxidation. Nature 2016, 530, 465–468. 10.1038/nature16529. [DOI] [PubMed] [Google Scholar]
  52. Chen Q.-F.; Cheng Z.-Y.; Liao R.-Z.; Zhang M.-T. Bioinspired Trinuclear Copper Catalyst for Water Oxidation with a Turnover Frequency up to 20000 s–1. J. Am. Chem. Soc. 2021, 143, 19761–19768. 10.1021/jacs.1c08078. [DOI] [PubMed] [Google Scholar]
  53. Lee K. J.; Elgrishi N.; Kandemir B.; Dempsey J. L. Electrochemical and spectroscopic methods for evaluating molecular electrocatalysts. Nat. Rev. Chem. 2017, 1, 0039 10.1038/s41570-017-0039. [DOI] [Google Scholar]
  54. Barnett S. M.; Goldberg K. I.; Mayer J. M. A soluble copper–bipyridine water-oxidation electrocatalyst. Nat. Chem. 2012, 4, 498–502. 10.1038/nchem.1350. [DOI] [PubMed] [Google Scholar]
  55. Kondo M.; Tatewaki H.; Masaoka S. Design of molecular water oxidation catalysts with earth-abundant metal ions. Chem. Soc. Rev. 2021, 50, 6790–6831. 10.1039/D0CS01442G. [DOI] [PubMed] [Google Scholar]
  56. Shaffer D. W.; Xie Y.; Szalda D. J.; Concepcion J. J. Lability and Basicity of Bipyridine-Carboxylate-Phosphonate Ligand Accelerate Single-Site Water Oxidation by Ruthenium-Based Molecular Catalysts. J. Am. Chem. Soc. 2017, 139, 15347–15355. 10.1021/jacs.7b06096. [DOI] [PubMed] [Google Scholar]
  57. Yang H.; Li F.; Zhan S.; Liu Y.; Li W.; Meng Q.; Kravchenko A.; Liu T.; Yang Y.; Fang Y.; Wang L.; Guan J.; Furó I.; Ahlquist M. S. G.; Sun L. Intramolecular hydroxyl nucleophilic attack pathway by a polymeric water oxidation catalyst with single cobalt sites. Nat. Catal. 2022, 5, 414–429. 10.1038/s41929-022-00783-6. [DOI] [Google Scholar]
  58. Hou S.; Xu L.; Ding X.; Kluge R. M.; Sarpey T. K.; Haid R. W.; Garlyyev B.; Mukherjee S.; Warnan J.; Koch M.; Zhang S.; Li W.; Bandarenka A. S.; Fischer R. A. Dual In Situ Laser Techniques Underpin the Role of Cations in Impacting Electrocatalysts. Angew. Chem., Int. Ed. 2022, 61, e202201610 10.1002/anie.202201610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. 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. 10.1039/C7CS00542C. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

Crystallographic data for the structures reported in this Article have been deposited at the Cambridge Crystallographic Data Centre under deposition numbers CCDC 2209276 and 2209277. Copies of the data can be obtained free of charge from www.ccdc.cam.ac.uk/structures/. Other data that support the findings of this study are available from the corresponding author upon reasonable request.


Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES