Skip to main content
JACS Au logoLink to JACS Au
. 2024 Mar 11;4(3):1207–1218. doi: 10.1021/jacsau.4c00047

Bioinspired Binickel Catalyst for Carbon Dioxide Reduction: The Importance of Metal–ligand Cooperation

Yao Xiao 1, Fei Xie 1, Hong-Tao Zhang 1, Ming-Tian Zhang 1,*
PMCID: PMC10976602  PMID: 38559717

Abstract

graphic file with name au4c00047_0008.jpg

Catalyst design for the efficient CO2 reduction reaction (CO2RR) remains a crucial challenge for the conversion of CO2 to fuels. Natural Ni–Fe carbon monoxide dehydrogenase (NiFe-CODH) achieves reversible conversion of CO2 and CO at nearly thermodynamic equilibrium potential, which provides a template for developing CO2RR catalysts. However, compared with the natural enzyme, most biomimetic synthetic Ni–Fe complexes exhibit negligible CO2RR catalytic activities, which emphasizes the significance of effective bimetallic cooperation for CO2 activation. Enlightened by bimetallic synergy, we herein report a dinickel complex, NiIINiII(bphpp)(AcO)2 (where NiNi(bphpp) is derived from H2bphpp = 2,9-bis(5-tert-butyl-2-hydroxy-3-pyridylphenyl)-1,10-phenanthroline) for electrocatalytic reduction of CO2 to CO, which exhibits a remarkable reactivity approximately 5 times higher than that of the mononuclear Ni catalyst. Electrochemical and computational studies have revealed that the redox-active phenanthroline moiety effectively modulates the electron injection and transfer akin to the [Fe3S4] cluster in NiFe-CODH, and the secondary Ni site facilitates the C–O bond activation and cleavage through electron mediation and Lewis acid characteristics. Our work underscores the significant role of bimetallic cooperation in CO2 reduction catalysis and provides valuable guidance for the rational design of CO2RR catalysts.

Keywords: CO2 reduction, binickel catalyst, homogeneous catalysis, C−O bond cleavage, bimetallic cooperative catalysis, redox-active ligand

Introduction

The reduction of CO2 to CO is a promising strategy for utilizing CO2 as a renewable C1 feedstock and establishing an artificial carbon cycle.15 Despite numerous studies on transition-metal-based CO2RR molecular catalysts, including those based on Re,610 Rh,1113 Ru,1423 Fe,2437 Ni,3844 Co,4555 Mn,5663 Cr,64,65 and so forth,6670 challenge persists in bridging the gap between current advances and practical applications, and thus the design of efficient molecular catalysts remains a significant challenge.7174 In nature, carbon monoxide dehydrogenase (CODH) plays a pivotal role in biological carbon fixation processes. This enzyme proceeds the binding of CO and CO2 to its Ni–Fe bimetallic active center, followed by electron transfer to the substrate, enabling the reversible conversion.7580 NiFe-CODH exemplifies the importance of bimetallic synergy in CO2 reduction catalysis: it activates CO2 at only −0.6 V versus standard hydrogen electrode (SHE) through the Ni site, with a [Fe3S4] cluster serving as an electron mediator. Upon protonation of Ni-CO2, the Fe site promotes the hydroxide elimination as a Lewis acid, thereby catalyzing the conversion of CO2 to CO (Scheme 1a).8183 This mechanism indicates that dinuclear metal complexes are promising candidates for CO2 reduction catalysis, as evidenced in several bimetallic catalytic systems.8499

Scheme 1. (a) Structure of the NiFe-CODH Active Site; (b) Proposed Two Modes of Bimetallic CO2 Activation; (c) CO2 Reduction by the Designed Dinickel Catalyst in This Work.

Scheme 1

Although structure similarities in bimetallic catalytic intermediates, such as Pd(μ22-CO2)Pd85,87 and Co(μ22-C,O:η1-O)Co,91 suggest that the structure feature of NiFe-CODH could inform the development of artificial bimetallic catalysts, the synthesized Ni–Fe clusters and macrocyclic Ni–Fe complexes with closely positioned metals have yet to demonstrate the catalytic activity for CO2 reduction.100109 These findings underscore that simply combining two metals does not replicate the catalytic function of NiFe-CODH and also highlight the necessity of integrating CO2 complexation, effective electron transfer, stabilization of intermediates, and an appropriate geometric arrangement of the catalyst to facilitate C–O bond cleavage and CO release. Meanwhile, two primary mechanisms of bimetallic cooperative CO2 activation have been proposed: M122-CO2)M2 and M12–C(O)OC(O)O−κ2-C,O)M2 (Scheme 1b).85,91,95,110,111 For the former, maintaining a close intermetallic distance (<3.0 Å) is crucial,108 which imposes a higher geometric constraint. Conversely, the latter mechanism can accommodate various geometries, potentially allowing for closer positioning and enhanced interaction between the metal sites.

Herein, as depicted in Scheme 1c, we designed a well-structured dinickel complex with a nearly planar geometry. Taking inspiration from the [Fe3S4] cluster in NiFe-CODH, we employed a redox-active phenanthroline ligand112 as an electron reservoir. This design facilitates electron transfer and enhances CO2 binding at the coordinated Ni1 site. Furthermore, the secondary Ni2 site, which possesses a distinct coordination environment, aids in the activation and cleavage of the C–O bond. This is achieved through a [Ni(μ2–C(O)OC(O)O−κ2-C,O)Ni] six-membered ring intermediate, which mitigates the geometry constraints associated with bimetallic centers. The binuclear [NiIINiII(bphpp)(AcO)2] complex (NiNi(bphpp)) demonstrates high reactivity in the catalytic reduction of CO2 to CO and promotes the generation of CO32– through ligand reduction and bimetallic synergy. This work holds the potential to make a substantial contribution to the development of novel strategies for CO2 capture and utilization.

Results and Discussion

Synthesis and Characterization of NiNi(bphpp) and Ni(hbpp)

H2bphpp (2,9-bis(5-tert-butyl-2-hydroxy-3-pyridylphenyl)-1,10-phenanthroline) and H2hbpp (2,9-bis(5-tert-butyl-2-hydroxyphenyl)-1,10-phenanthroline) are synthesized by the modified procedures113 and used as the dinuclear and mononuclear ligands, respectively (Scheme S1, Figures S1–S15). The dinickel complex NiNi(bphpp) and mononickel complex [NiII(hbpp)] (Ni(hbpp)) are prepared by mixing the corresponding ligand with 2 equiv or 1 equiv Ni(AcO)2·4H2O, respectively, which are detailed in the Supporting Information (SI). The X-ray crystal structure of NiNi(bphpp) shows that the two NiII ions are both in an octahedral geometry configuration, which are bridged by two phenolic O atoms and connected by two acetates located above and below the ligand plane (Figures 1a and S16). The distance between the two NiII sites is 2.70 Å, which is similar to the distance of 2.8 Å between Ni and Fe in CODH,81,83,108 indicating a suitable geometric structure required for CO2 activation. In contrast, the X-ray crystal structure of Ni(hbpp) shows that the NiII ion has a square-planar configuration (Figures 1b and S17), consistent with low-spin NiII (S = 0) characterized by 1H NMR (Figure S18). NiNi(bphpp) and Ni(hbpp) were also characterized by high-resolution mass spectrometry (HRMS) in methanol, which showed the molecular cation peak at m/z 372.0763 ([NiIINiII(bphpp)]2+, Figure S19) and m/z 555.1557 ([NiII(hbpp) + Na]+, Figure S20), respectively. These results suggest that NiNi(bphpp) and Ni(hbpp) exist as di- and mononickel species in solution, respectively. In addition, the density functional theory (DFT) study showed that the dissociation energy of acetate in NiNi(bphpp) is 13.5 kcal/mol (Figure S41), indicating the consistency of the crystal structure with that in solution.

Figure 1.

Figure 1

Schematic structures and X-ray crystal structures of NiNi(bphpp) (a), (b) and Ni(hbpp) (c), (d), respectively. Ellipsoids are shown at 80% probability of atoms and 60% probability of bonds. Two acetate anions are drawn in wireframe style; the hydrogen atoms and solvent molecules are omitted for clarity.

The electronic configuration of NiNi(bphpp) is further investigated. The Evans effective magnetic moment (μeff) was measured to be 4.16 μB at 290 K (Table S4, detailed in SI),114,115 which is consistent with the spin multiplicity of S = 2. The magnetic properties of NiNi(bphpp) were obtained over the temperature range of 1.8–300 K by a superconducting quantum interference device (SQUID, Figure S21). The calculated effective magnetic moment, μeff, was 4.5 μB at the range 200–300 K, which is consistent with the measurement obtained using the Evans method, indicating a total spin state of S = 2. As the temperature decreases, the magnetic susceptibility increases, reaching a maximum of 5.3 μB at 10 K, which is close to the expected μeff for a two-spin-only S = 1 NiII system (μeff.s.o = 2.8 × 2 = 5.6 μB), suggesting the two NiII ions are weakly ferromagnetic coupled. These results agree well with the DFT calculation of the ground state electronic structure of NiNi(bphpp) that indicates a quintuple state (Figure S48). The μeff. decreases to 4.0 μB at 1.8 K, which could be ascribed to the zero-field splitting (zfs) of the ground state. The two NiII ions of NiNi(bphpp) are assumed to be equivalent with the same zfs parameters D1, rhombicity E/D, and electronic g-tensors. VT-VH magnetic susceptibility was collected at 1, 4, and 7 T (Figure S21). Global simulation of the VT-VH magnetic susceptibility and magnetization data, based on the spin-Hamiltonian (Section 1 of SI) of the low-lying electronic states, yielded an exchange coupling J = 7.4 cm–1, and D1 = D2 = −9.4 cm–1 with strongly anisotropic g values of gx = gy = 2.10 and gz = 2.33 for each NiII ion. The X-band EPR spectra of NiNi(bphpp) for orientations of the microwave field B1 parallel and perpendicular to static field H were both obtained at 5 K in frozen DMF solution (Figure S22), and the valleys at g = 12.6 and g = 9.9 were observed, respectively. For this non-Kramers dinuclear Ni system, large D1 values of each NiII ion were obtained from the magnetic data fitting and comparable to the exchange coupling, J, causing the S = 2 and S = 1 multiplets to mix,116118 and the zero-field absorption would origin from the energy levels with splitting in energy Δ ≈ 0.3 cm–1. Simulation of the parallel mode and perpendicular mode spectra gives E/D ≈ 0.12 with Dstrain = 0.1 cm–1 and Estrain = 0.05 cm–1 for the broad lines.

Electrochemical Studies of NiNi(bphpp) and Ni(hbpp)

The redox behavior of NiNi(bphpp) was examined by cyclic voltammogram (CV) and differential pulse voltammetry (DPV) measurements in anhydrous degassed DMF. The CV and DPV of NiNi(bphpp) show four consecutive reduction waves at −1.68, – 1.80, – 2.14, and −2.44 V vs Fc+/0 (Figures 2a and S25, all potentials below quoted vs Fc+/0). The scan-rate-dependent CVs illustrate that the first wave is reversible with a peak-to-peak separation (ΔEp) of 70 mV at a scan rate of 0.1 V/s (Figure S26), suggestive of a one-electron reduction process. The subsequent three waves are found to be irreversible (Figure S25). To elucidate the electrochemical processes associated with each reduction peak, a trizinc complex, Zn3(bphpp)(AcO)4, was synthesized (Figures S23 and S24). Its CV displayed two successive irreversible reduction waves at −1.86 and −1.98 V (Figure S27),119 indicating that the initial two reduction peaks in the CV of NiNi(bphpp) are attributable to the reduction of the phenanthroline moiety. DFT studies were employed to further investigate the electrochemical processes. The one-electron reduction of NiNi(bphpp) (1) leads to the formation of the NiIINiII(bphpp–•)(AcO)2 species (2), which exhibits a spin population of 0.82 on the phenanthroline moiety, and the potential of this reduction process is calculated to be −1.65 V (eq 1, Figure S49). At the calculated potential of −1.69 V, NiIINiII(bphpp–•)(AcO)2 can accept an additional electron at the phenanthroline moiety to generate NiIINiII(bphpp2–)(AcO)2 (3+2AcO, Figure S50). The two AcO ligands of 3+2AcO readily dissociate with a Gibbs free energy change of −11.3 kcal/mol, which is followed by charge redistribution, leading to the formation of NiINiI(bphpp) (3, eq 2). Moreover, even at a scan rate of 2.5 V/s, the second reduction wave remains irreversible, suggesting that the dissociation rate of the two AcO groups exceeds a certain threshold (Figure S28). Additionally, the reverse oxidation scan following the second reduction wave is tentatively ascribed to two-electron oxidation of the NiINiI(bphpp) species. Subsequently, the phenanthroline unit of NiINiI(bphpp) undergoes two successive one-electron reductions at −1.75 and −2.27 V, resulting in the generations of NiINiI(bphpp–•) (4, eq 3, Figure S51) and NiINiI(bphpp2–) (4red, eq 4, Figure S52), respectively. These findings align with the observed experimental electrochemical behavior and highlight the role of the phenanthroline moiety as an electron reservoir. For Ni(hbpp), the CV and DPV show two reduction peaks at −1.83 and −2.46 V, which are more negative than those observed for NiNi(bphpp) (Figures 2a and S29). The scan-rate-dependent CVs for Ni(hbpp) also indicate that the first wave is reversible, with a ΔEp value of 66 mV at 0.1 V/s (Figure S30), consistent with a one-electron reduction. These reduction potentials are similar to those reported for (tbuthbpy)FeCl by Machan et al.,120 which are proposed to represent two consecutive one-electron reductions, leading to the formation of NiII(hbpp–•) and NiII(hbpp2–), respectively. These findings are consistent with DFT results (see Figure S60). The CV comparisons demonstrate that both the phenanthroline moiety and the secondary Ni site in NiNi(bphpp) could help mitigate charge accumulation.

graphic file with name au4c00047_0011.jpg 1
graphic file with name au4c00047_0012.jpg 2
graphic file with name au4c00047_0013.jpg 3
graphic file with name au4c00047_0014.jpg 4

Figure 2.

Figure 2

(a) CVs of 0.5 mM NiNi(bphpp) and 0.5 mM Ni(hbpp) under Ar atmosphere; (b) CVs of 0.5 mM NiNi(bphpp) (red line) and 0.5 mM Ni(hbpp) (blue line) under saturated CO2 atmosphere, and CV under saturated CO2 without catalyst (black line) using a glassy carbon electrode at 100 mV/s in 0.1 M nBu4NPF6/DMF electrolyte.

In CO2-saturated anhydrous DMF solution (0.23 M4), the CV of NiNi(bphpp) shows a significant current enhancement, with a half peak potential (Ecat/2) of −2.25 V and peak potential at −2.40 V (Figure 2b).121 The overpotential is determined to be 1.52 V concerning the estimated potential for the CO2/CO couple in anhydrous DMF. this value is comparable to those reported for past catalytic systems in aprotic solutions.60,95,122,123 In contrast, the Ecat/2 of Ni(hbpp) is ca. −2.43 V under the same condition (Figure 2b), where the overpotential is more negative by 0.18 V than that of the dinickel catalyst. This suggests that the introduction of a secondary Ni site decreases the overpotential for catalysis by facilitating charge accumulation. According to the CVs of NiNi(bphpp) under Ar and CO2 atmosphere, the third irreversible reduction wave is attributed to the catalytic CO2 reduction by the NiINiI(bphpp–•) species. For Ni(hbpp), the electrochemical properties indicate that the species formed after two-electron reduction, that is NiII(hbpp2–) (Figure S60), is responsible for catalyzing CO2 reduction.

To confirm the products of electrocatalytic CO2 reduction, controlled potential electrolysis (CPE) of 0.5 mM NiNi(bphpp) at −2.20 V was carried out using a foamed graphite electrode (1.0 × 1.0 × 0.6 cm3). The electrolysis current remained stable at an absolute value of approximately 2.8 mA, with a total charge consumption of 53.6 coulombs over 5 h (Figures 3a and S31). Gas chromatography (GC) and ion chromatography (IC) were used to detect the products in the gas phase above the electrolysis cell and the electrolyte solution, respectively. The major products were identified as CO and CO32– (Figures S32 and S33). The CO production was quantified at 250 μmol over 5 h by GC (Figure 3b), while the turnover number (TON) and the Faradaic Efficiency (FE) were calculated to be 50.2 and 91%, respectively (Table S5). Furthermore, the production of CO32– was determined to be 267 μmol with a FE of 97% (Figure S33a). The measured quantity of CO32– is marginally higher than that of CO, which can be attributed to the inherent deviation in the quantification of CO32–.6 Additionally, negligible amounts of H2 and HCOO were detected (Figures S32 and S34), which was presumed to be due to the trace amounts of H2O in DMF. Gas chromatography–mass spectrometry (GC–MS) of CPE under a 13CO2 atmosphere confirmed the formation of 13CO (Figure S35). These results indicate that the reductive disproportionation of CO2 occurs without the presence of a Brönsted acid. In contrast, CPE of 0.5 mM Ni(hbpp) under the same conditions resulted in an electrolysis current of only about 0.5 mA, and product analysis indicated the formation of merely 45 μmol CO and 49 μmol CO32– (Figures 3b and S33b) over 5 h CPE with FEs of CO and CO32– are 90% and 99%, respectively. These results illustrate that Ni(hbpp) is less active toward the electrocatalytic reduction of CO2 compared to NiNi(bphpp), suggesting that the secondary Ni site plays an important role in the catalytic process of CO2 reduction disproportionation.

Figure 3.

Figure 3

(a) CPE current curves of 0.5 mM NiNi(bphpp) (red line), 0.5 mM Ni(hbpp) (blue line) in anhydrous 0.1 M nBu4NPF6/DMF, and blank solution (gray line) at −2.20 V; (b) CO production curves of 0.5 mM NiNi(bphpp) (red dot), 0.5 mM Ni(hbpp) (blue dot), and no catalyst (gray dot) during CPE at −2.20 V in 0.1 M nBu4NPF6/DMF CO2-saturated electrolyte.

Confirmation of Homogeneous Catalysis

To confirm the homogeneous nature of the catalysis, stability tests of the catalysts were conducted to determine whether (a) the catalyst gradually decomposes throughout the CPE; and (b) the active heterogeneous materials form and attach to the electrode surface. Due to the large current when using the foamed graphite electrode, it is difficult to observe the difference before and after electrolysis. Consequently, a plate glassy carbon electrode was chosen as the working electrode. After 4 h CPE at −2.20 V, CV of 0.1 M nBu4NPF6/DMF blank solution without either NiNi(bphpp) or Ni(hbpp) showed no obvious current change by using the plate glassy carbon electrode after rinse (Figure S36). Additionally, CVs of 0.5 mM NiNi(bphpp) or Ni(hbpp) solution after 4 h of CPE with the rinsed electrode displayed only minor differences compared to the CVs obtained before electrolysis with a fresh electrode (Figure S36). Further examination of the surface of the electrolyzed electrode through X-ray photoelectron spectroscopy (XPS) and scanning electron microscopy (SEM) revealed that the morphology and composition of rinsed plate glassy carbon electrode remained consistent with an unused electrode, with no visible heterogeneous materials detected (Figures S37 and S38). These findings strongly support the assertion that the catalysis processes involving NiNi(bphpp) and Ni(hbpp) are stable and homogeneous.

Kinetic Analysis

To intuitively assess the catalytic efficiency of NiNi(bphpp) and Ni(hbpp), the foot of wave analysis (FOWA) method is applied to determine the turnover frequency (TOF).124126 Our results show that the calculated TOF of NiNi(bphpp), derived from catalytic CV was 20.5 s–1, while that of Ni(hbpp) was determined to be 4.4 s–1 (Table S6, Figures S39 and S40).127 The improved TOFcalc obtained by the introduction of the secondary Ni site demonstrates the significant enhancement of catalytic activity through bimetallic cooperation.

Furthermore, the kinetics of electrocatalytic CO2 reduction of NiNi(bphpp) and Ni(hbpp) were explored to understand the specific role of the secondary Ni site. The catalytic peak current (icat) of NiNi(bphpp) showed a linear relationship with its concentration (Figure S41), indicating the first-order kinetics of NiNi(bphpp). Similarly, a linear relationship was observed between icat and the concentration of Ni(hbpp) (Figure S42). These findings suggest that both complexes facilitate CO2 reduction as single molecules within the concentration range of 0–1.0 mM. However, the kinetic orders of CO2 in the catalytic processes of the two catalysts differed markedly. Within the saturated range of CO2 in DMF, the catalytic process of NiNi(bphpp) exhibited two distinct kinetic order regimes (Figures 4a and S43). At CO2 concertation ([CO2]) below 0.04 M, icat was directly proportional to the concentration of CO2, indicating second-order kinetics. In contrast, icat correlated with the square root of CO2 concentration ([CO2]1/2) between 0.04 and 0.23 M (Inset of Figure 4a), suggesting first-order kinetics. In comparison, in the saturated range of CO2, the catalytic process of Ni(hbpp) followed zero-order kinetics (Figures 4b and S44). Considering the three key steps in the catalytic process of CO2 reduction disproportionation (i.e., CO2 binding, C–O bond activation, and C–O bond cleavage, eqs S5–S7),8,111,123 it appears that NiNi(bphpp) presents a relatively low barrier of C–O bond cleavage; the CO2 binding and C–O bond activation processes collectively dictate the catalytic rate at lower CO2 concentration (eq S12), while C–O bond activation, which has a higher barrier, becomes more dominant as the concentration of CO2 increases (eq S13). Conversely, the zero-order kinetics for Ni(hbpp) concerning [CO2] implies an extraordinarily high barrier of C–O bond cleavage, as CO2 is not involved in this step (eq S14). Integrating with DFT study (discussed below and in Section 5 of the SI), the kinetic results strongly indicate that the secondary Ni site plays a private role in facilitating C–O bond activation and cleavage, thereby lowering the barriers in the catalytic cycle and potentially alerting the rate-determining step of the reaction. The augmented catalytic capability conferred by the well-integrated secondary Ni site was corroborated by the combined product analysis and the TOF and TON values. Additionally, a mixing experiment was conducted for further validation. Mixing Ni(hbpp) with an equivalent of Ni(AcO)2 under identical electrochemical conditions does not result in any significant enhancement of the catalytic current (Figure S45).

Figure 4.

Figure 4

Relationship between icat and [CO2] of NiNi(bphpp) (a) and Ni(hbpp) (b) in the [CO2] range of 0–0.23 M. Inset: the relationship between icat and [CO2]1/2 of NiNi(bphpp) for the one-order area in (a).

Computational Studies

We further investigated the role of the secondary Ni site in enhancing the catalytic reduction of CO2 using DFT calculations. The Gibbs free energy diagram (Figure S46) illustrates that CO2 binding and subsequent chemical steps are facilitated following the formation of NiINiI(bphpp–•) species 4 (Figure 5a). Under our inference, CO2 binds readily to the Ni1 site, which is coordinated by reduced phenanthroline, thereby exhibiting increased nucleophilic character. This significant property is demonstrated by the low energy barrier associated with the reaction. Specifically, the Ni1–C1 bond forms via transition state TS1, with a remarkably low barrier of 3.8 kcal/mol, while the formation of the Ni2–C1 bond has a considerably higher barrier of 14.8 kcal/mol relative to species 4 (Figure S52), resulting in the generation of Int1 (Figures 5b and S53). Interestingly, the Ni-CO2 species within Int1 presents a unique geometry, with a C1–O1 and C1–O2 distance of 1.21 Å, a Ni1–C1 distance of 2.094 Å, and an O1–C1–O2 angle of 146.9°. These parameters differ markedly from the previously reported structures of NiII1-CO22–-κC species which exhibit longer C–O distance (1.24–1.25 Å), shorter Ni–C distance (1.91–1.95 Å), and smaller angle of O–C–O (128–129°).130,131 Hence, Int1 represents a transitional configuration between NiII–CO2 and NiII–CO22–, presumably a NiII–CO2–• species. Moreover, during the formation of Int1, the spin population on Ni1 increases from 0.98 to 1.17, and the overall spin population on CO2 increases to −0.363. This antiferromagnetic coupling with Ni1 gives rise to the distinct NiINiII(bphpp–•)(CO2–•) species. Subsequently, Int1 can attract another CO2 molecule due to the impending nucleophilic attack of the O1 on C2, which carries an emerging negative charge, leading to the generation of intermediate Int2 (Figures 5c and S53), where the O1–C2 distance is 2.79 Å.

Figure 5.

Figure 5

Spin population analysis (isovalue of 0.01) of 4 (a), Int1 (b), Int2 (c), TS2 (d), and Int3 (e) (Phen = phenanthroline) by Multiwfn128 and VMD.129 Green and blue isosurfaces represent α and β electrons, respectively. Spin population values are shown in italics. (f) The bimetallic cooperation mode of C–O bond activation and cleavage regulated by NiNi(bphpp).

Int2 subsequently initiates the critical C2–O1 bond formation between the CO2–• moiety and the second CO2 with the notable assistance of Ni2. Within this cooperative context, the incoming CO2 molecule can readily approach the reduced CO2–• and activate the C1–O1 bond through transition state TS2 (Figures 5d and S54). The spin population of Ni1 increases from 1.21 to 1.59, while the spin population of phenanthroline decreases from 0.74 to 0.10, indicating an electron transfer from the phenanthroline ligand to the CO2–• moiety. Consequently, the resulting CO22– unit can nucleophilically attack the adjacent CO2 molecule with a relatively low barrier of 12.4 kcal/mol to Int2 (Figure 6). Following this cooperative binding of activated CO2 with additional CO2, the OCOCO22– intermediate (Int3) is formed (Figures 5e and S55). Notably, the spin population of Ni2 increases from 0.97 in TS2 to 1.73, whereas that of phenanthroline reverts to 0.56. This marked change in spin populations indicates an electron transfer from the Ni2(I) site to the phenanthroline, resulting in the formation of Ni2(II), which serves to stabilize the OCOCO22– anion with its Lewis acid character and culminates in the formation of the corresponding NiIINiII(bphpp–•)(OCOCO22–) species (Int3). In addition, the C1–O1 bond length gradually increases from 1.21 Å (Int1) to 1.22 Å (Int2), and finally to 1.44 Å (Int3), indicative of a typical C–O single bond. This bond activation process comprehensively corroborates the essential cooperative functions of the secondary Ni site (Figure 5f), which (a) acts as an electronic reservoir by storing the reduction equivalents, (b) facilitates the spatial approach of another CO2 molecule as an electron mediator, thereby enabling electron transfers from reductive phenanthroline moiety to CO2–• moiety, and also from Ni2 to phenanthroline, and (c) stabilizes the OCOCO22– anion as a Lewis acid site.

Figure 6.

Figure 6

Calculated Gibbs free-energy diagrams of NiNi(bphpp) (blue line) and Ni(hbpp) (black line) starting at Int1 and Int1S at the B3LYP-D3/6-311+G(2df,2p)||B3LYP-D3/6-31G(d,p) level.

Subsequently, the OCOCO22– unit within Int3 can be converted into more thermodynamically stable species CO and CO32– due to the cleavage of the C1–O1 bond. The CO–CO32– cleavage occurs via a highly favorable transition state TS3 (Figure S56), which is characterized by a comparatively lower barrier of 8.3 kcal/mol. This leads to the formation of a transient intermediate, termed Int4 (Figure S57). The spin populations on Ni1, Ni2, and phenanthroline (1.26, 1.69, and 0.46, respectively) remain relatively unchanged from those of Int3, indicating a direct C–O cleavage process that results in the formation of the NiIINiII(bphpp–•)(CO)(CO32–) complex. This facile cleavage of the C–O bond is attributed to the Lewis acid character of Ni2, which also stabilizes the CO32– unit (Figure 5f). Two pathways for the release of CO and CO32– were considered, and both are considered plausible (Figure S46). Initially, the direct release pathway of either CO32– or CO is endergonic by 17 or 0.6 kcal/mol, leading to the generation of Int5 or Int5′ (Figure S58), respectively. However, when coupled with a one-electron reduction, both processes become exergonic by −21.8 and −8 kcal/mol, resulting in the formation of Int6 and Int6 (Figure S59), respectively. Although the prior release of CO32– after a one-electron reduction is thermodynamically favorable, the subsequent reductive release of CO to complete the catalytic cycle is challenging due to the reduction potential of −2.12 V. In contrast, the reductive release of CO32– from Int6 is facilitated by a lower potential of −1.52 V and returns to species 4.

The importance of the secondary Ni site is further illustrated by the catalytic process of Ni(hbpp), denoted as 1S. In contrast to 1, the phenanthroline in Ni(hbpp) undergoes two consecutive one-electron reductions, resulting in the generations of 2S and 3S, respectively (Figure S60). The spin population of phenanthroline rises from 0 to 0.79 and then to 1.51, while that of Ni remains almost unchanged. The binding of CO2 occurs via transition state TS1S after the formation of 3S with a low barrier of 2.2 kcal/mol (Figures S47 and S61). Similar to NiNi(bphpp), this process involves the one-electron activation of CO2 with the generation of NiII(hbpp–•)(CO2–•), labeled as Int1S, which has a relatively high reduction potential of −2.56 V (Figure S62). Thus, we inferred that Int1S directly forms the C2–O1 bond with the second CO2 molecule via transition state TS2S (Figures 6 and S63), resulting in the generation of the five-membered ring intermediate, Int2S (Figure S64). Without the cooperation of the secondary Ni site, the barrier of this bonding process is calculated to be 18.7 kcal/mol, which is 2.8 kcal/mol higher than that of NiNi(bphpp) (Figure 6). Subsequently, Int2S undergoes C1–O1 cleavage via transition state TS3S (Figure S65), followed by the departure of CO to generate NiII(hbpp)(CO32–) intermediate Int3S (Figure S66), which is exergonic by 23 kcal/mol (Figures 6 and S47). The barrier calculated for TS3S is as high as 21.3 kcal/mol, which shows an obvious difference from the dinickel catalyst. For the transformation of Int3 to Int4 via TS3, Ni1(II) coordinated with reductive phenanthroline exhibits a certain nucleophilic character, which stabilizes the CO unit during C–O cleavage. In addition, Ni2(II) also stabilizes the CO32– unit as a Lewis acid. This bimetallic cooperation results in a much lower barrier for the formation of Int4. However, compared with the only C–O bond cleavage in TS3, TS3S involves both C–O bond cleavage and CO dissociation on a single NiII site, which is undoubtedly more kinetically unfavorable. The high C–O bond cleavage barrier for TS3S also supports the zero-order kinetics of the catalytic process toward [CO2]. Furthermore, the energetic span model (detailed in SI) is utilized to better describe the overall catalytic rate through the Gibbs free energy diagrams (Figures S46 and 47).50,132,133 According to the model equation eq S15, the overall kinetic barrier (δEspan) for the NiNi(bphpp) catalytic cycle is calculated to be 19.3 kcal/mol, which is significantly lower than the δEspan value of 22.1 kcal/mol for the catalytic cycle of Ni(hbpp). Using eq S16, the calculated TOF for NiNi(bphpp) catalysis is expected to be significantly higher than that of Ni(hbpp) catalysis. These theoretical results are completely consistent with the electrochemical properties, product characterizations, and catalytic kinetics of NiNi(bphpp) and Ni(hbpp).

Both controlled experiments and DFT calculations have substantiated the pronounced enhancement of redox-active phenanthroline in facilitating CO2 activation through coordinated Ni site, which avoids the generation of high-energy Ni0 and contributes to the one-electron reduction of CO2 to CO2–•. Furthermore, the distinct roles of both Ni sites have been elucidated. Notably, the secondary Ni site actively aids in the activation of the C–O bond through “pre-coordination” with another CO2 molecule as an important electron mediator. Additionally, the Lewis acid character of the secondary Ni site markedly facilitates the C–O bond cleavage, thereby promoting the entire catalytic process. This work emphasizes that bimetallic cooperation and redox-active ligand assistance are effective design strategies for the development of CO2 reduction catalysts, even for the multielectron/proton redox catalysis of the other small molecules.134141

Conclusions

Inspired by the catalytic mechanism of NiFe-CODH, this work explores the development of a binuclear Ni catalyst, NiNi(bphpp), which enables bimetallic synergistic catalysis for CO2 activation and catalytic conversion. By incorporating redox-active phenanthroline and a secondary Ni site, the bimetallic NiNi(bphpp) catalyst enables a three-electron reduction at a lower potential, leading to the generation of NiINiI(bphpp–•) species. Subsequently, facile one-electron activation of CO2 occurs, followed by activation and cleavage of the C–O bond through the coordination of a second CO2 molecule. The involvement of the secondary Ni(I) site acts as an effective electron mediator, promoting electron transfer to the second CO2 molecule and facilitating C–O bond activation. Furthermore, after formation of the [OCOCO22–] species, the secondary Ni(II) site exhibits Lewis acid characteristics, significantly aiding in C–O bond cleavage. These findings shed light on the distinct roles played by bimetallic sites in CO2 reduction catalysis, providing insights into the nature of bimetallic cooperation. This work proposes new design strategies for catalysts in bimetallic synergistic multielectron catalytic processes. Based on this strategy, the development of the Ni–Fe heterobimetallic catalyst is in progress for efficient CO2 reduction.

Methods

All of the electrochemical experiments were conducted with a three-electrode system using a CHI-660E electrochemical workstation at room temperature. A glassy carbon electrode (GCE, 0.07 cm2) was used as the working electrode, the counter electrode was a Pt wire, and Ag/AgNO3 (0.1 M AgNO3 in DMF) was used as the reference electrode and calibrated versus Fc+/0. All of the reduction potentials were obtained from the DPVs of corresponding complexes. The gas above the electrolytic cell was drawn by a 1 mL Hamilton syringe and injected into the GC. CPE was conducted with 10 mL of 0.1 M TBAPF6/DMF solution containing NiNi(bphpp) or Ni(hbpp) (0.5 mM) in a sealed standard three-electrode cell with stirring. A foamy carbon electrode (1.0 × 1.0 × 0.6 cm) was used as the working electrode. The Faraday efficiencies were calculated by the equation below:

graphic file with name au4c00047_m001.jpg
graphic file with name au4c00047_m002.jpg

where COexp represents the CO content above the electrolytic cell which was determined by gas chromatography, and COCPE represents the CO content calculated from the total charge transfer (Q) during CPE (Figure S31).

Acknowledgments

We acknowledge the National Natural Science Foundation of China (NSFC 21933007 and 22193011) for funding support. We also appreciate Dr. Qiang Han for the support in ion chromatography analysis.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.4c00047.

  • Additional experimental details, materials, methods, complex synthesis (1H NMR and ESI-HRMS), magnetic property characterizations (Evans magnetic moment, EPR, and SQUID), X-ray crystallography, electrochemical studies, kinetic analysis, DFT calculations, and so on (PDF)

Accession Codes

CCDC 2257038, 2257041, 2257045, and 2257048 contain supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: + 44 1223 336033.

The authors declare no competing financial interest.

Supplementary Material

au4c00047_si_001.pdf (19MB, pdf)

References

  1. Lewis N. S.; Nocera D. G. Powering the planet: chemical challenges in solar energy utilization. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 15729–15735. 10.1073/pnas.0603395103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Benson E. E.; Kubiak C. P.; Sathrum A. J.; Smieja J. M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. 10.1039/B804323J. [DOI] [PubMed] [Google Scholar]
  3. Appel A. M.; Bercaw J. E.; Bocarsly A. B.; Dobbek H.; DuBois D. L.; Dupuis M.; Ferry J. G.; Fujita E.; Hille R.; Kenis P. J. A.; Kerfeld C. A.; Morris R. H.; Peden C. H.; Portis A. R.; Ragsdale S. W.; Rauchfuss T. B.; Reek J. N. H.; Seefeldt L. C.; Thauer R. K.; Waldrop G. L. Frontiers, opportunities, and challenges in biochemical and chemical catalysis of CO2 fixation. Chem. Rev. 2013, 113, 6621–6658. 10.1021/cr300463y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Francke R.; Schille B.; Roemelt M. Homogeneously catalyzed electroreduction of carbon dioxide-methods, mechanisms, and catalysts. Chem. Rev. 2018, 118, 4631–4701. 10.1021/acs.chemrev.7b00459. [DOI] [PubMed] [Google Scholar]
  5. De Luna P.; Hahn C.; Higgins D.; Jaffer S. A.; Jaramillo T. F.; Sargent E. H. What would it take for renewably powered electrosynthesis to displace petrochemical processes?. Science 2019, 364, eaav3506 10.1126/science.aav3506. [DOI] [PubMed] [Google Scholar]
  6. Sullivan B. P.; Bolinger C. M.; Conrad D.; Vining W. J.; Meyer T. J. One- and two-electron pathways in the electrocatalytic reduction of CO2 by fac-Re(bpy)(CO)3Cl (bpy = 2,2’-bipyridine). J. Chem. Soc., Chem. Commun. 1985, 1414–1416. 10.1039/C39850001414. [DOI] [Google Scholar]
  7. Keith J. A.; Grice K. A.; Kubiak C. P.; Carter E. A. Elucidation of the selectivity of proton-dependent electrocatalytic CO2 reduction by fac-Re(bpy)(CO)3Cl. J. Am. Chem. Soc. 2013, 135, 15823–15829. 10.1021/ja406456g. [DOI] [PubMed] [Google Scholar]
  8. Machan C. W.; Chabolla S. A.; Yin J.; Gilson M. K.; Tezcan F. A.; Kubiak C. P. Supramolecular assembly promotes the electrocatalytic reduction of carbon dioxide by Re(I) bipyridine catalysts at a lower overpotential. J. Am. Chem. Soc. 2014, 136, 14598–14607. 10.1021/ja5085282. [DOI] [PubMed] [Google Scholar]
  9. Sung S.; Kumar D.; Gil-Sepulcre M.; Nippe M. Electrocatalytic CO2 reduction by imidazolium-functionalized molecular catalysts. J. Am. Chem. Soc. 2017, 139, 13993–13996. 10.1021/jacs.7b07709. [DOI] [PubMed] [Google Scholar]
  10. Haviv E.; Azaiza-Dabbah D.; Carmieli R.; Avram L.; Martin J. M. L.; Neumann R. A thiourea tether in the second coordination sphere as a binding site for CO2 and a proton donor promotes the electrochemical reduction of CO2 to CO catalyzed by a rhenium bipyridine-type complex. J. Am. Chem. Soc. 2018, 140, 12451–12456. 10.1021/jacs.8b05658. [DOI] [PubMed] [Google Scholar]
  11. Slater S.; Wagenknecht J. H. Electrochemical reduction of carbon dioxide catalyzed by Rh(diphos)2Cl. J. Am. Chem. Soc. 1984, 106, 5367–5368. 10.1021/ja00330a064. [DOI] [Google Scholar]
  12. Bolinger C. M.; Story N.; Sullivan B. P.; Meyer T. J. Electrocatalytic reduction of carbon dioxide by 2,2’-bipyridine complexes of rhodium and iridium. Inorg. Chem. 1988, 27, 4582–4587. 10.1021/ic00298a016. [DOI] [Google Scholar]
  13. Todorova T. K.; Huan T. N.; Wang X.; Agarwala H.; Fontecave M. Controlling hydrogen evolution during photoreduction of CO2 to formic acid using [Rh(R-bpy)(Cp*)Cl]+ catalysts: a structure-activity study. Inorg. Chem. 2019, 58, 6893–6903. 10.1021/acs.inorgchem.9b00371. [DOI] [PubMed] [Google Scholar]
  14. Ishida H.; Tanaka H.; Tanaka K.; Tanaka T. Selective formation of HCOO in the electrochemical CO2 reduction catalysed by [Ru(bpy)2(CO)2]2+ (bpy = 2,2′-bipyridine). J. Chem. Soc., Chem. Commun. 1987, 131–132. 10.1039/C39870000131. [DOI] [Google Scholar]
  15. Mizukawa T.; Tsuge K.; Nakajima H.; Tanaka K. Selective production of acetone in the electrochemical reduction of CO2 catalyzed by a Ru-naphthyridine complex. Angew. Chem., Int. Ed. 1999, 38, 362–363. . [DOI] [PubMed] [Google Scholar]
  16. Boston D. J.; Pachón Y. M.; Lezna R. O.; de Tacconi N. R.; MacDonnell F. M. Electrocatalytic and photocatalytic conversion of CO2 to methanol using ruthenium complexes with internal pyridyl cocatalysts. Inorg. Chem. 2014, 53, 6544–6553. 10.1021/ic500051m. [DOI] [PubMed] [Google Scholar]
  17. Machan C. W.; Sampson M. D.; Kubiak C. P. A molecular ruthenium electrocatalyst for the reduction of carbon dioxide to CO and formate. J. Am. Chem. Soc. 2015, 137, 8564–8571. 10.1021/jacs.5b03913. [DOI] [PubMed] [Google Scholar]
  18. Johnson B. A.; Maji S.; Agarwala H.; White T. A.; Mijangos E.; Ott S. Activating a low overpotential CO2 reduction mechanism by a strategic ligand modification on a ruthenium polypyridyl catalyst. Angew. Chem., Int. Ed. 2016, 55, 1825–1829. 10.1002/anie.201508490. [DOI] [PubMed] [Google Scholar]
  19. Ono T.; Qu S.; Gimbert-Suriñach C.; Johnson M. A.; Marell D. J.; Benet-Buchholz J.; Cramer C. J.; Llobet A. Hydrogenative carbon dioxide reduction catalyzed by mononuclear ruthenium polypyridyl complexes: discerning between electronic and steric effects. ACS Catal. 2017, 7, 5932–5940. 10.1021/acscatal.7b00603. [DOI] [Google Scholar]
  20. Gonell S.; Massey M. D.; Moseley I. P.; Schauer C. K.; Muckerman J. T.; Miller A. J. M. The trans effect in electrocatalytic CO2 reduction: mechanistic studies of asymmetric ruthenium pyridyl-carbene catalysts. J. Am. Chem. Soc. 2019, 141, 6658–6671. 10.1021/jacs.9b01735. [DOI] [PubMed] [Google Scholar]
  21. Sampaio R. N.; Grills D. C.; Polyansky D. E.; Szalda D. J.; Fujita E. Unexpected roles of triethanolamine in the photochemical reduction of CO2 to formate by ruthenium complexes. J. Am. Chem. Soc. 2020, 142, 2413–2428. 10.1021/jacs.9b11897. [DOI] [PubMed] [Google Scholar]
  22. Agarwala H.; Chen X.; Lyonnet J. R.; Johnson B. A.; Ahlquist M.; Ott S. Alternating metal-ligand coordination improves electrocatalytic CO2 reduction by a mononuclear Ru catalyst. Angew. Chem., Int. Ed. 2023, 62, e202218728 10.1002/anie.202218728. [DOI] [PubMed] [Google Scholar]
  23. Hunt L. A.; Das S.; Lamb R. W.; Nugegoda D.; Curiac C.; Figgins M. T.; Lambert E. C.; Qu F.; Hammer N. I.; Delcamp J. H.; Webster C. E.; Papish E. T. Ruthenium (II) complexes of CNC pincers and bipyridine in the photocatalytic CO2 reduction reaction to CO using visible light: catalysis, kinetics, and computational insights. ACS Catal. 2023, 13, 5986–5999. 10.1021/acscatal.2c05459. [DOI] [Google Scholar]
  24. Hammouche M.; Lexa D.; Savéant J. M.; Momenteau M. Catalysis of the electrochemical reduction of carbon dioxide by iron(“0”) porphyrins. J. Electroanal. Chem. 1988, 249, 347–351. 10.1016/0022-0728(88)80372-3. [DOI] [Google Scholar]
  25. Costentin C.; Drouet S.; Robert M.; Savéant J.-M. A local proton source enhances CO2 electroreduction to CO by a molecular Fe catalyst. Science 2012, 338, 90–94. 10.1126/science.1224581. [DOI] [PubMed] [Google Scholar]
  26. Chen L.; Guo Z.; Wei X. G.; Gallenkamp C.; Bonin J.; Anxolabéhère-Mallart E.; Lau K.-C.; Lau T.-C.; Robert M. Molecular catalysis of the electrochemical and photochemical reduction of CO2 with earth-abundant metal complexes. selective production of CO vs HCOOH by switching of the metal center. J. Am. Chem. Soc. 2015, 137, 10918–10921. 10.1021/jacs.5b06535. [DOI] [PubMed] [Google Scholar]
  27. Azcarate I.; Costentin C.; Robert M.; Savéant J. M. Through-space charge interaction substituent effects in molecular catalysis leading to the design of the most efficient catalyst of CO2-to-CO electrochemical conversion. J. Am. Chem. Soc. 2016, 138, 16639–16644. 10.1021/jacs.6b07014. [DOI] [PubMed] [Google Scholar]
  28. Rao H.; Schmidt L. C.; Bonin J.; Robert M. Visible-light-driven methane formation from CO2 with a molecular iron catalyst. Nature 2017, 548, 74–77. 10.1038/nature23016. [DOI] [PubMed] [Google Scholar]
  29. Margarit C. G.; Schnedermann C.; Asimow N. G.; Nocera D. G. Carbon dioxide reduction by iron hangman porphyrins. Organometallics 2019, 38, 1219–1223. 10.1021/acs.organomet.8b00334. [DOI] [Google Scholar]
  30. Derrick J. S.; Loipersberger M.; Chatterjee R.; Iovan D. A.; Smith P. T.; Chakarawet K.; Yano J.; Long J. R.; Head-Gordon M.; Chang C. J. Metal–ligand cooperativity via exchange coupling promotes iron-catalyzed electrochemical CO2 reduction at low overpotentials. J. Am. Chem. Soc. 2020, 142, 20489–20501. 10.1021/jacs.0c10664. [DOI] [PubMed] [Google Scholar]
  31. Martin D. J.; Mayer J. M. Oriented electrostatic effects on O2 and CO2 reduction by a polycationic iron porphyrin. J. Am. Chem. Soc. 2021, 143, 11423–11434. 10.1021/jacs.1c03132. [DOI] [PubMed] [Google Scholar]
  32. Wang X.-Z.; Meng S.-L.; Chen J.-Y.; Wang H.-X.; Wang Y.; Zhou S.; Li X.-B.; Liao R.-Z.; Tung C.-H.; Wu L.-Z. Mechanistic insights into iron(II) bis(pyridyl)amine-bipyridine skeleton for selective CO2 photoreduction. Angew. Chem., Int. Ed. 2021, 60, 26072–26079. 10.1002/anie.202107386. [DOI] [PubMed] [Google Scholar]
  33. Amanullah S.; Saha P.; Dey A. Activating the Fe(I) state of iron porphyrinoid with second-sphere proton transfer residues for selective reduction of CO2 to HCOOH via Fe(III/II)-COOH intermediate(s). J. Am. Chem. Soc. 2021, 143, 13579–13592. 10.1021/jacs.1c04392. [DOI] [PubMed] [Google Scholar]
  34. Massie A. A.; Schremmer C.; Rüter I.; Dechert S.; Siewert I.; Meyer F. Selective electrocatalytic CO2 reduction to CO by an NHC-based organometallic heme analogue. ACS Catal. 2021, 11, 3257–3267. 10.1021/acscatal.0c04518. [DOI] [Google Scholar]
  35. Guo K.; Li X.; Lei H.; Guo H.; Jin X.; Zhang X.-P.; Zhang W.; Apfel U.-P.; Cao R. Role-specialized division of labor in CO2 reduction with doubly-functionalized iron porphyrin atropisomers. Angew. Chem., Int. Ed. 2022, 61, e202209602 10.1002/anie.202209602. [DOI] [PubMed] [Google Scholar]
  36. Derrick J. S.; Loipersberger M.; Nistanaki S. K.; Rothweiler A. V.; Head-Gordon M.; Nichols E. M.; Chang C. J. Templating bicarbonate in the second coordination sphere enhances electrochemical CO2 reduction catalyzed by iron porphyrins. J. Am. Chem. Soc. 2022, 144, 11656–11663. 10.1021/jacs.2c02972. [DOI] [PubMed] [Google Scholar]
  37. Pugliese E.; Gotico P.; Wehrung I.; Boitrel B.; Quaranta A.; Ha-Thi M. H.; Pino T.; Sircoglou M.; Leibl W.; Halime Z.; Aukauloo A. Dissection of light-induced charge accumulation at a highly active iron porphyrin: insights in the photocatalytic CO2 reduction. Angew. Chem., Int. Ed. 2022, 61, e202117530 10.1002/anie.202117530. [DOI] [PubMed] [Google Scholar]
  38. Fisher B.; Eisenberg R. Electrocatalytic reduction of carbon dioxide by using macrocycles of nickel and cobalt. J. Am. Chem. Soc. 1980, 102, 7361–7363. 10.1021/ja00544a035. [DOI] [Google Scholar]
  39. Beley M.; Collin J. P.; Ruppert R.; Sauvage J. P. Electrocatalytic reduction of CO2 by Ni cyclam2+ in water: study of the factors affecting the efficiency and the selectivity of the process. J. Am. Chem. Soc. 1986, 108, 7461–7467. 10.1021/ja00284a003. [DOI] [PubMed] [Google Scholar]
  40. Rudolph M.; Dautz S.; Jäger E.-G. Macrocyclic [N42-] coordinated nickel complexes as catalysts for the formation of oxalate by electrochemical reduction of carbon dioxide. J. Am. Chem. Soc. 2000, 122, 10821–10830. 10.1021/ja001254n. [DOI] [Google Scholar]
  41. Thoi V. S.; Kornienko N.; Margarit C. G.; Yang P.; Chang C. J. Visible-light photoredox catalysis: selective reduction of carbon dioxide to carbon monoxide by a nickel N-heterocyclic carbene-isoquinoline complex. J. Am. Chem. Soc. 2013, 135, 14413–14424. 10.1021/ja4074003. [DOI] [PubMed] [Google Scholar]
  42. Froehlich J. D.; Kubiak C. P. The homogeneous reduction of CO2 by [Ni(cyclam)]+: increased catalytic rates with the addition of a CO scavenger. J. Am. Chem. Soc. 2015, 137, 3565–3573. 10.1021/ja512575v. [DOI] [PubMed] [Google Scholar]
  43. Hong D.; Tsukakoshi Y.; Kotani H.; Ishizuka T.; Kojima T. Visible-light-driven photocatalytic CO2 reduction by a Ni(II) complex bearing a bioinspired tetradentate ligand for selective CO production. J. Am. Chem. Soc. 2017, 139, 6538–6541. 10.1021/jacs.7b01956. [DOI] [PubMed] [Google Scholar]
  44. Shirley H.; Su X.; Sanjanwala H.; Talukdar K.; Jurss J. W.; Delcamp J. H. Durable solar-powered systems with Ni-catalysts for conversion of CO2 or CO to CH4. J. Am. Chem. Soc. 2019, 141, 6617–6622. 10.1021/jacs.9b00937. [DOI] [PubMed] [Google Scholar]
  45. Lacy D. C.; McCrory C. C.; Peters J. C. Studies of cobalt-mediated electrocatalytic CO2 reduction using a redox-active ligand. Inorg. Chem. 2014, 53, 4980–4988. 10.1021/ic403122j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Shaffer D. W.; Johnson S. I.; Rheingold A. L.; Ziller J. W.; Goddard W. A. III; Nielsen R. J.; Yang J. Y. Reactivity of a series of isostructural cobalt pincer complexes with CO2, CO, and H+. Inorg. Chem. 2014, 53, 13031–13041. 10.1021/ic5021725. [DOI] [PubMed] [Google Scholar]
  47. Chapovetsky A.; Do T. H.; Haiges R.; Takase M. K.; Marinescu S. C. Proton-assisted reduction of CO2 by cobalt aminopyridine macrocycles. J. Am. Chem. Soc. 2016, 138, 5765–5768. 10.1021/jacs.6b01980. [DOI] [PubMed] [Google Scholar]
  48. Guo Z.; Cheng S.; Cometto C.; Anxolabéhère-Mallart E.; Ng S.-M.; Ko C.-C.; Liu G.; Chen L.; Robert M.; Lau T.-C. Highly efficient and selective photocatalytic CO2 Reduction by iron and cobalt quaterpyridine complexes. J. Am. Chem. Soc. 2016, 138, 9413–9416. 10.1021/jacs.6b06002. [DOI] [PubMed] [Google Scholar]
  49. Roy S.; Sharma B.; Pecaut J.; Simon P.; Fontecave M.; Tran P. D.; Derat E.; Artero V. Molecular cobalt complexes with pendant amines for selective electrocatalytic reduction of carbon dioxide to formic acid. J. Am. Chem. Soc. 2017, 139, 3685–3696. 10.1021/jacs.6b11474. [DOI] [PubMed] [Google Scholar]
  50. Fernández S.; Franco F.; Casadevall C.; Martin-Diaconescu V.; Luis J. M.; Lloret-Fillol J. A unified electro- and photocatalytic CO2 to CO reduction mechanism with aminopyridine cobalt complexes. J. Am. Chem. Soc. 2020, 142, 120–133. 10.1021/jacs.9b06633. [DOI] [PubMed] [Google Scholar]
  51. Dey S.; Todorova T. K.; Fontecave M.; Mougel V. Electroreduction of CO2 to formate with low overpotential using cobalt pyridine thiolate complexes. Angew. Chem., Int. Ed. 2020, 59, 15726–15733. 10.1002/anie.202006269. [DOI] [PubMed] [Google Scholar]
  52. Queyriaux N.; Abel K.; Fize J.; Pécaut J.; Orio M.; Hammarström L. From non-innocent to guilty: on the role of redox-active ligands in the electro-assisted reduction of CO2 mediated by a cobalt(II)-polypyridyl complex. Sustainable Energy Fuels 2020, 4, 3668–3676. 10.1039/D0SE00570C. [DOI] [Google Scholar]
  53. Zhang X.; Yamauchi K.; Sakai K. Earth-abundant photocatalytic CO2 reduction by multielectron chargeable cobalt porphyrin catalysts: high CO/H2 selectivity in water based on phase mismatch in frontier MO association. ACS Catal. 2021, 11, 10436–10449. 10.1021/acscatal.1c02475. [DOI] [Google Scholar]
  54. Loipersberger M.; Cabral D. G. A.; Chu D. B. K.; Head-Gordon M. Mechanistic insights into Co and Fe quaterpyridine-based CO2 reduction catalysts: metal-ligand orbital interaction as the key driving force for distinct pathways. J. Am. Chem. Soc. 2021, 143, 744–763. 10.1021/jacs.0c09380. [DOI] [PubMed] [Google Scholar]
  55. Wang X. S.; Yang J. Y. Translating aqueous CO2 hydrogenation activity to electrocatalytic reduction with a homogeneous cobalt catalyst. Chem. Commun. 2023, 59, 338–341. 10.1039/D2CC05473F. [DOI] [PubMed] [Google Scholar]
  56. Bourrez M.; Molton F.; Chardon-Noblat S.; Deronzier A. [Mn(bipyridyl)(CO)3Br]: an abundant metal carbonyl complex as efficient electrocatalyst for CO2 reduction. Angew. Chem., Int. Ed. 2011, 50, 9903–9906. 10.1002/anie.201103616. [DOI] [PubMed] [Google Scholar]
  57. Sampson M. D.; Nguyen A. D.; Grice K. A.; Moore C. E.; Rheingold A. L.; Kubiak C. P. Manganese catalysts with bulky bipyridine ligands for the electrocatalytic reduction of carbon dioxide: eliminating dimerization and altering catalysis. J. Am. Chem. Soc. 2014, 136, 5460–5471. 10.1021/ja501252f. [DOI] [PubMed] [Google Scholar]
  58. Agarwal J.; Shaw T. W.; Stanton C. J. III; Majetich G. F.; Bocarsly A. B.; Schaefer H. F. III NHC-containing manganese(I) electrocatalysts for the two-electron reduction of CO2. Angew. Chem., Int. Ed. 2014, 53, 5152–5155. 10.1002/anie.201311099. [DOI] [PubMed] [Google Scholar]
  59. Ngo K. T.; McKinnon M.; Mahanti B.; Narayanan R.; Grills D. C.; Ertem M. Z.; Rochford J. Turning on the protonation-first pathway for electrocatalytic CO2 reduction by manganese bipyridyl tricarbonyl complexes. J. Am. Chem. Soc. 2017, 139, 2604–2618. 10.1021/jacs.6b08776. [DOI] [PubMed] [Google Scholar]
  60. Franco F.; Pinto M. F.; Royo B.; Lloret-Fillol J. A highly active N-heterocyclic carbene manganese(I) complex for selective electrocatalytic CO2 reduction to CO. Angew. Chem., Int. Ed. 2018, 57, 4603–4606. 10.1002/anie.201800705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Sung S.; Li X.; Wolf L. M.; Meeder J. R.; Bhuvanesh N. S.; Grice K. A.; Panetier J. A.; Nippe M. Synergistic effects of imidazolium-functionalization on fac-Mn(CO)3 bipyridine catalyst platforms for electrocatalytic carbon dioxide reduction. J. Am. Chem. Soc. 2019, 141, 6569–6582. 10.1021/jacs.8b13657. [DOI] [PubMed] [Google Scholar]
  62. Ro̷nne M. H.; Cho D.; Madsen M. R.; Jakobsen J. B.; Eom S.; Escoudé É.; Hammersho̷j H. C. D.; Nielsen D. U.; Pedersen S. U.; Baik M.-H.; Skrydstrup T.; Daasbjerg K. Ligand-controlled product selectivity in electrochemical carbon dioxide reduction using manganese bipyridine catalysts. J. Am. Chem. Soc. 2020, 142, 4265–4275. 10.1021/jacs.9b11806. [DOI] [PubMed] [Google Scholar]
  63. Yang Y.; Ertem M. Z.; Duan L. An amide-based second coordination sphere promotes the dimer pathway of Mn-catalyzed CO2-to-CO reduction at low overpotential. Chem. Sci. 2021, 12, 4779–4788. 10.1039/D0SC05679K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Tory J.; Setterfield-Price B.; Dryfe R. A. W.; Hartl F. [M(CO)4 (2,2’-bipyridine)] (M = Cr, Mo, W) complexes as efficient catalysts for electrochemical reduction of CO2 at a gold electrode. ChemElectroChem. 2015, 2, 213–217. 10.1002/celc.201402282. [DOI] [Google Scholar]
  65. Hooe S. L.; Moreno J. J.; Reid A. G.; Cook E. N.; Machan C. W. Mediated inner-sphere electron transfer induces homogeneous reduction of CO2 via through-space electronic conjugation. Angew. Chem., Int. Ed. 2022, 61, e202109645 10.1002/anie.202205139. [DOI] [PubMed] [Google Scholar]
  66. DuBois D. L.; Miedaner A.; Haltiwanger R. C. Electrochemical reduction of carbon dioxide catalyzed by [Pd(triphosphine)(solvent)](BF4)2 complexes: synthetic and mechanistic studies. J. Am. Chem. Soc. 1991, 113, 8753–8764. 10.1021/ja00023a023. [DOI] [Google Scholar]
  67. Therrien J. A.; Wolf M. O.; Patrick B. O. Electrocatalytic reduction of CO2 with palladium bis-N-heterocyclic carbene pincer complexes. Inorg. Chem. 2014, 53, 12962–12972. 10.1021/ic502056w. [DOI] [PubMed] [Google Scholar]
  68. Buss J. A.; VanderVelde D. G.; Agapie T. Lewis acid enhancement of proton induced CO2 cleavage: Bond weakening and ligand residence time effects. J. Am. Chem. Soc. 2018, 140, 10121–10125. 10.1021/jacs.8b05874. [DOI] [PubMed] [Google Scholar]
  69. Cunningham D. W.; Barlow J. M.; Velazquez R. S.; Yang J. Y. Reversible and selective CO2 to HCO2 electrocatalysis near the thermodynamic potential. Angew. Chem., Int. Ed. 2020, 59, 4443–4447. 10.1002/anie.201913198. [DOI] [PubMed] [Google Scholar]
  70. Ceballos B. M.; Yang J. Y. Highly selective electrocatalytic CO2 reduction by [Pt(dmpe)2]2+ through kinetic and thermodynamic control. Organometallics 2020, 39, 1491–1496. 10.1021/acs.organomet.9b00720. [DOI] [Google Scholar]
  71. Costentin C.; Robert M.; Savéant J. M. Current issues in molecular catalysis illustrated by iron porphyrins as catalysts of the CO2-to-CO electrochemical conversion. Acc. Chem. Res. 2015, 48, 2996–3006. 10.1021/acs.accounts.5b00262. [DOI] [PubMed] [Google Scholar]
  72. Amanullah S.; Saha P.; Nayek A.; Ahmed M. E.; Dey A. Biochemical and artificial pathways for the reduction of carbon dioxide, nitrite and the competing proton reduction: effect of 2nd sphere interactions in catalysis. Chem. Soc. Rev. 2021, 50, 3755–3823. 10.1039/D0CS01405B. [DOI] [PubMed] [Google Scholar]
  73. Fujita E.; Grills D. C.; Manbeck G. F.; Polyansky D. E. Understanding the role of inter- and intramolecular promoters in electro- and photochemical CO2 reduction using Mn, Re, and Ru catalysts. Acc. Chem. Res. 2022, 55, 616–628. 10.1021/acs.accounts.1c00616. [DOI] [PubMed] [Google Scholar]
  74. Saha P.; Amanullah S.; Dey A. Selectivity in electrochemical CO2 reduction. Acc. Chem. Res. 2022, 55, 134–144. 10.1021/acs.accounts.1c00678. [DOI] [PubMed] [Google Scholar]
  75. Shin W.; Lindahl P. A. Function and CO binding properties of the NiFe complex in carbon monoxide dehydrogenase from Clostridium thermoaceticum. Biochemistry 1992, 31, 12870–12875. 10.1021/bi00166a023. [DOI] [PubMed] [Google Scholar]
  76. Ragsdale S. W.; Kumar M. Nickel-containing carbon monoxide dehydrogenase/acetyl-CoA synthase. Chem. Rev. 1996, 96, 2515–2539. 10.1021/cr950058+. [DOI] [PubMed] [Google Scholar]
  77. Gong W.; Hao B.; Wei Z.; Ferguson D. J. Jr.; Tallant T.; Krzycki J. A.; Chan M. K. Structure of the α2ε2 Ni-dependent CO dehydrogenase component of the Methanosarcina barkeri acetyl-CoA decarbonylase/synthase complex. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 9558–9563. 10.1073/pnas.0800415105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Can M.; Armstrong F. A.; Ragsdale S. W. Structure, function, and mechanism of the nickel metalloenzymes, CO dehydrogenase, and acetyl-CoA synthase. Chem. Rev. 2014, 114, 4149–4174. 10.1021/cr400461p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Stripp S. T.; Duffus B. R.; Fourmond V.; Léger C.; Leimkühler S.; Hirota S.; Hu Y.; Jasniewski A.; Ogata H.; Ribbe M. W. Second and outer coordination sphere effects in nitrogenase, hydrogenase, formate dehydrogenase, and CO dehydrogenase. Chem. Rev. 2022, 122, 11900–11973. 10.1021/acs.chemrev.1c00914. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Basak Y.; Jeoung J. H.; Domnik L.; Dobbek H. Stepwise O2-induced rearrangement and disassembly of the [NiFe4(OH)(μ3-S)4] active site cluster of CO dehydrogenase. Angew. Chem., Int. Ed. 2023, 62, e202305341 10.1002/anie.202305341. [DOI] [PubMed] [Google Scholar]
  81. Dobbek H.; Svetlitchnyi V.; Gremer L.; Huber R.; Meyer O. Crystal structure of a carbon monoxide dehydrogenase reveals a [Ni-4Fe-5S] cluster. Science 2001, 293, 1281–1285. 10.1126/science.1061500. [DOI] [PubMed] [Google Scholar]
  82. Dobbek H.; Svetlitchnyi V.; Liss J.; Meyer O. Carbon monoxide induced decomposition of the active site [Ni-4Fe-5S] cluster of CO dehydrogenase. J. Am. Chem. Soc. 2004, 126, 5382–5387. 10.1021/ja037776v. [DOI] [PubMed] [Google Scholar]
  83. Jeoung J. H.; Dobbek H. Carbon dioxide activation at the Ni,Fe-cluster of anaerobic carbon monoxide dehydrogenase. Science 2007, 318, 1461–1464. 10.1126/science.1148481. [DOI] [PubMed] [Google Scholar]
  84. Haines R. J.; Wittrig R. E.; Kubiak C. P. Electrocatalytic reduction of carbon dioxide by the binuclear copper complex [Cu2(6-(diphenylphosphino)-2,2’-bipyridyl)2(MeCN)2][PF6]2. Inorg. Chem. 1994, 33, 4723–4728. 10.1021/ic00099a024. [DOI] [Google Scholar]
  85. Steffey B. D.; Curtis C. J.; Dubois D. L. Electrochemical reduction of CO2 catalyzed by a dinuclear palladium complex containing a bridging hexaphosphine ligand: evidence for cooperativity. Organometallics 1995, 14, 4937–4943. 10.1021/om00010a066. [DOI] [Google Scholar]
  86. Ali M. M.; Sato H.; Mizukawa T.; Tsuge K.; Haga M.; Tanaka K. Selective formation of HCO2 and C2O42- in electrochemical reduction of CO2 catalyzed by mono- and di-nuclear ruthenium complexes. Chem. Commun. 1998, 249–250. 10.1039/a707363a. [DOI] [Google Scholar]
  87. Raebiger J. W.; Turner J. W.; Noll B. C.; Curtis C. J.; Miedaner A.; Cox B.; DuBois D. L. Electrochemical reduction of CO2 to CO catalyzed by a bimetallic palladium complex. Organometallics 2006, 25, 3345–3351. 10.1021/om060228g. [DOI] [Google Scholar]
  88. Angamuthu R.; Byers P.; Lutz M.; Spek A. L.; Bouwman E. Electrocatalytic CO2 conversion to oxalate by a copper complex. Science 2010, 327, 313–315. 10.1126/science.1177981. [DOI] [PubMed] [Google Scholar]
  89. Mohamed E. A.; Zahran Z. N.; Naruta Y. Efficient electrocatalytic CO2 reduction with a molecular cofacial iron porphyrin dimer. Chem. Commun. 2015, 51, 16900–16903. 10.1039/C5CC04273A. [DOI] [PubMed] [Google Scholar]
  90. Witt S. E.; White T. A.; Li Z.; Dunbar K. R.; Turro C. Cationic dirhodium(II,II) complexes for the electrocatalytic reduction of CO2 to HCOOH. Chem. Commun. 2016, 52, 12175–12178. 10.1039/C6CC03253B. [DOI] [PubMed] [Google Scholar]
  91. Ouyang T.; Huang H. H.; Wang J. W.; Zhong D. C.; Lu T. B. A dinuclear cobalt cryptate as a homogeneous photocatalyst for highly selective and efficient visible-light driven CO2 reduction to CO in CH3CN/H2O solution. Angew. Chem., Int. Ed. 2017, 56, 738–743. 10.1002/anie.201610607. [DOI] [PubMed] [Google Scholar]
  92. Cao L.-M.; Huang H.-H.; Wang J.-W.; Zhong D.-C.; Lu T.-B. The synergistic catalysis effect within a dinuclear nickel complex for efficient and selective electrocatalytic reduction of CO2 to CO. Green Chem. 2018, 20, 798–803. 10.1039/C7GC03451B. [DOI] [Google Scholar]
  93. Cheng M.; Yu Y.; Zhou X.; Luo Y.; Wang M. Chemical versatility of [FeFe]-hydrogenase models: distinctive activity of [μ-C6H4-1,2-(κ2-S)2][Fe2(CO)6] for electrocatalytic CO2 reduction. ACS Catal. 2019, 9, 768–774. 10.1021/acscatal.8b03921. [DOI] [Google Scholar]
  94. Ouyang T.; Wang H.-J.; Huang H.-H.; Wang J.-W.; Guo S.; Liu W.-J.; Zhong D.-C.; Lu T.-B. Dinuclear metal synergistic catalysis boosts photochemical CO2-to-CO conversion. Angew. Chem., Int. Ed. 2018, 57, 16480–16485. 10.1002/anie.201811010. [DOI] [PubMed] [Google Scholar]
  95. Yang W.; Sinha Roy S.; Pitts W. C.; Nelson R. L.; Fronczek F. R.; Jurss J. W. Electrocatalytic CO2 reduction with cis and trans conformers of a rigid dinuclear rhenium complex: Comparing the monometallic and cooperative bimetallic pathways. Inorg. Chem. 2018, 57, 9564–9575. 10.1021/acs.inorgchem.8b01775. [DOI] [PubMed] [Google Scholar]
  96. Guo Z.; Chen G.; Cometto C.; Ma B.; Zhao H.; Groizard T.; Chen L.; Fan H.; Man W.-L.; Yiu S.-M.; Lau K.-C.; Lau T.-C.; Robert M. Selectivity control of CO versus HCOO production in the visible-light-driven catalytic reduction of CO2 with two cooperative metal sites. Nat. Catal. 2019, 2, 801–808. 10.1038/s41929-019-0331-6. [DOI] [Google Scholar]
  97. Ahmed M. E.; Adam S.; Saha D.; Fize J.; Artero V.; Dey A.; Duboc C. Repurposing a bio-inspired NiFe hydrogenase model for CO2 reduction with selective production of methane as the unique C-based product. ACSEnergy Lett. 2020, 5, 3837–3842. 10.1021/acsenergylett.0c02002. [DOI] [Google Scholar]
  98. Polyansky D. E.; Grills D. C.; Ertem M. Z.; Ngo K. T.; Fujita E. Role of bimetallic interactions in the enhancement of catalytic CO2 reduction by a macrocyclic cobalt catalyst. ACS Catal. 2022, 12, 1706–1717. 10.1021/acscatal.1c05043. [DOI] [Google Scholar]
  99. Zhang C.; Gotico P.; Guillot R.; Dragoe D.; Leibl W.; Halime Z.; Aukauloo A. Bio-inspired bimetallic cooperativity through a hydrogen bonding spacer in CO2 reduction. Angew. Chem., Int. Ed. 2023, 62, e202214665 10.1002/anie.202214665. [DOI] [PubMed] [Google Scholar]
  100. Ciurli S.; Ross P. K.; Scott M. J.; Yu S.-B.; Holm R. H. Synthetic nickel-containing heterometal cubane-type clusters with NiFe3Q4 cores (Q = S, Se). J. Am. Chem. Soc. 1992, 114, 5415–5423. 10.1021/ja00039a063. [DOI] [Google Scholar]
  101. Tan G. O.; Ensign S. A.; Ciurli S.; Scott M. J.; Hedman B.; Holm R. H.; Ludden P. W.; Korszun Z. R.; Stephens P. J.; Hodgson K. O. On the structure of the nickel/iron/sulfur center of the carbon monoxide dehydrogenase from Rhodospirillum rubrum: an X-ray absorption spectroscopy study. Proc. Natl. Acad. Sci. U. S. A. 1992, 89, 4427–4431. 10.1073/pnas.89.10.4427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Zhou J.; Scott M. J.; Hu Z.; Peng G.; Munck E.; Holm R. H. Synthesis and comparative reactivity and electronic structural features of [MFe3S4]z+ cubane-type clusters (M = Fe, Co, Ni). J. Am. Chem. Soc. 1992, 114, 10843–10854. 10.1021/ja00053a021. [DOI] [Google Scholar]
  103. Osterloh F.; Saak W.; Pohl S. Unidentate and bidentate binding of nickel(II) complexes to an Fe4S4 cluster via bridging thiolates: Synthesis, crystal structures, and electrochemical properties of model compounds for the active sites of nickel containing CO dehydrogenase/Acetyl-CoA synthase. J. Am. Chem. Soc. 1997, 119, 5648–5656. 10.1021/ja970194r. [DOI] [Google Scholar]
  104. Panda R.; Zhang Y.; McLauchlan C. C.; Rao P. V.; Tiago de Oliveira F. A.; Munck E.; Holm R. H. Initial structure modification of tetrahedral to planar nickel(II) in a nickel-iron-sulfur cluster related to the C-cluster of carbon monoxide dehydrogenase. J. Am. Chem. Soc. 2004, 126, 6448–6459. 10.1021/ja030627s. [DOI] [PubMed] [Google Scholar]
  105. Sun J.; Tessier C.; Holm R. H. Sulfur ligand substitution at the nickel(II) sites of cubane-type and cubanoid NiFe3S4 clusters relevant to the C-clusters of carbon monoxide dehydrogenase. Inorg. Chem. 2007, 46, 2691–2699. 10.1021/ic062362z. [DOI] [PubMed] [Google Scholar]
  106. Majumdar A. Bioinorganic modeling chemistry of carbon monoxide dehydrogenases: description of model complexes, current status and possible future scopes. Dalton Trans. 2014, 43, 12135–12145. 10.1039/C4DT00729H. [DOI] [PubMed] [Google Scholar]
  107. Fesseler J.; Jeoung J. H.; Dobbek H. How the [NiFe4S4] cluster of CO dehydrogenase activates CO2 and NCO. Angew. Chem., Int. Ed. 2015, 54, 8560–8564. 10.1002/anie.201501778. [DOI] [PubMed] [Google Scholar]
  108. Huang D.; Holm R. H. Reactions of the terminal NiII-OH group in substitution and electrophilic reactions with carbon dioxide and other substrates: structural definition of binding modes in an intramolecular NiII···FeII bridged site. J. Am. Chem. Soc. 2010, 132, 4693–4701. 10.1021/ja1003125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Prat J. R.; Gaggioli C. A.; Cammarota R. C.; Bill E.; Gagliardi L.; Lu C. C. Bioinspired nickel complexes supported by an iron metalloligand. Inorg. Chem. 2020, 59, 14251–14262. 10.1021/acs.inorgchem.0c02041. [DOI] [PubMed] [Google Scholar]
  110. Ratliff K. S.; Lentz R. E.; Kubiak C. P. Carbon dioxide chemistry of the trinuclear complex [Ni33-CNMe)(μ3-I)(dppm)3][PF6]. Electrocatalytic reduction of carbon dioxide. Organometallics 1992, 11, 1986–1988. 10.1021/om00042a004. [DOI] [Google Scholar]
  111. Agarwal J.; Fujita E.; Schaefer H. F. III; Muckerman J. T. Mechanisms for CO production from CO2 using reduced rhenium tricarbonyl catalysts. J. Am. Chem. Soc. 2012, 134, 5180–5186. 10.1021/ja2105834. [DOI] [PubMed] [Google Scholar]
  112. Pun S.-N.; Chung W.-H.; Lam K.-M.; Guo P.; Chan P.-H.; Wong K.-Y.; Che C.-M.; Chen T.-Y.; Peng S.-M. Iron(I) complexes of 2,9-bis(2-hydroxyphenyl)-1,10-phenanthroline (H2dophen) as electrocatalysts for carbon dioxide reduction. X-ray crystal structures of [Fe(dophen)Cl]2·2HCON(CH3)2 and [Fe(dophen)(N-MeIm)2]ClO4 (N-MeIm = 1-methylimidazole). J. Chem. Soc., Dalton Trans. 2002, 575–583. 10.1039/b108472k. [DOI] [Google Scholar]
  113. Lam F.; Feng M.; Chan K. S. Synthesis of dinucleating phenanthroline-based ligands. Tetrahedron 1999, 55, 8377–8384. 10.1016/S0040-4020(99)00428-7. [DOI] [Google Scholar]
  114. Evans D. F. The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance. J. Chem. Soc. 1959, 1959, 2003–2005. 10.1039/jr9590002003. [DOI] [Google Scholar]
  115. Löliger J.; Scheffold R. Paramagnetic moment measurements by NMR. a micro technique. J. Chem. Educ. 1972, 49, 646–647. 10.1021/ed049p646. [DOI] [Google Scholar]
  116. Juarez-Garcia C.; Hendrich M. P.; Holman T. R.; Que L.; Munck E. Combined Möessbauer and EPR studies of the S = 3 state of an exchange-coupled FeIIICuII complex: test for quantitative EPR analysis of integer spin systems. J. Am. Chem. Soc. 1991, 113, 518–525. 10.1021/ja00002a020. [DOI] [Google Scholar]
  117. Hendrich M. P.; Day E. P.; Wang C.-P.; Synder B. S.; Holm R. H.; Münck E. Ground-state electronic structures of binuclear iron(II) sites: experimental protocol and a consistent description of Möessbauer, EPR, and magnetization measurements of the bis(phenolate)-bridged complex [Fe2(salmp)2]2-. Inorg. Chem. 1994, 33, 2848–2856. 10.1021/ic00091a028. [DOI] [Google Scholar]
  118. Reger D. L.; Pascui A. E.; Smith M. D.; Jezierska J.; Ozarowski A. Dinuclear complexes containing linear M–F–M [M = Mn(II), Fe(II), Co(II), Ni(II), Cu(II), Zn(II), Cd(II)] bridges: trends in structures, antiferromagnetic superexchange interactions, and spectroscopic properties. Inorg. Chem. 2012, 51, 11820–11836. 10.1021/ic301757g. [DOI] [PubMed] [Google Scholar]
  119. The structure and overall charge of Zn3(bphpp)(AcO)4 are different from those of NiNi(bphpp). The CV comparison of Zn3(bphpp)(AcO)4 and NiNi(bphpp) is only intended to be utilized to assign the reduction processes of the phenanthroline unit.
  120. Nichols A. W.; Chatterjee S.; Sabat M.; Machan C. W. Electrocatalytic reduction of CO2 to formate by an iron schiff base complex. Inorg. Chem. 2018, 57, 2111–2121. 10.1021/acs.inorgchem.7b02955. [DOI] [PubMed] [Google Scholar]
  121. 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]
  122. Machan C. W.; Chabolla S. A.; Kubiak C. P. Reductive Disproportionation of Carbon Dioxide by an Alkyl-Functionalized Pyridine Monoimine Re(I) fac-Tricarbonyl Electrocatalyst. Organometallics 2015, 34, 4678–4683. 10.1021/acs.organomet.5b00406. [DOI] [Google Scholar]
  123. Sampson M. D.; Kubiak C. P. Manganese electrocatalysts with bulky bipyridine ligands: utilizing Lewis acids to promote carbon dioxide reduction at low overpotentials. J. Am. Chem. Soc. 2016, 138, 1386–1393. 10.1021/jacs.5b12215. [DOI] [PubMed] [Google Scholar]
  124. Costentin C.; Drouet S.; Robert M.; Savéant J.-M. Turnover numbers, turnover frequencies, and overpotential in molecular catalysis of electrochemical reactions. Cyclic voltammetry and preparative-scale electrolysis. J. Am. Chem. Soc. 2012, 134, 11235–11242. 10.1021/ja303560c. [DOI] [PubMed] [Google Scholar]
  125. Costentin C.; Savéant J.-M. Multielectron, multistep molecular catalysis of electrochemical reactions: benchmarking of homogeneous catalysts. ChemElectroChem. 2014, 1, 1226–1236. 10.1002/celc.201300263. [DOI] [Google Scholar]
  126. Wang V. C.-C.; Johnson B. A. Interpreting the electrocatalytic voltammetry of homogeneous catalysts by the foot of the wave analysis and its wider implications. ACS Catal. 2019, 9, 7109–7123. 10.1021/acscatal.9b00850. [DOI] [Google Scholar]
  127. Since the electrocatalytic CO2 reduction disproportionation at corresponding reduction peak is an electrochemical-multi-chemical process, the TOFcalc is the rate of the chemical rate-determining step (C-O activation step of NiNi(bphpp) and C-O cleavage step of Ni(hbpp)) following the reduction but not the TOFmax.
  128. Lu T.; Chen F. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. 10.1002/jcc.22885. [DOI] [PubMed] [Google Scholar]
  129. Humphrey W.; Dalke A.; Schulten K. VMD: visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
  130. Yoo C.; Lee Y. Carbon dioxide binding at a Ni/Fe center: synthesis and characterization of Ni(η1-CO2C) and Ni-μ-CO2C2O. O’-Fe. Chem. Sci. 2017, 8, 600–605. 10.1039/C6SC03450K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Oren D.; Diskin-Posner Y.; Avram L.; Feller M.; Milstein D. Metal-ligand cooperation as key in formation of dearomatized NiII-H pincer complexes and in their reactivity toward CO and CO2. Organometallics 2018, 37, 2217–2221. 10.1021/acs.organomet.8b00160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Kozuch S.; Shaik S. How to conceptualize catalytic cycles? The energetic span model. Acc. Chem. Res. 2011, 44, 101–110. 10.1021/ar1000956. [DOI] [PubMed] [Google Scholar]
  133. Garay-Ruiz D.; Bo C. Revisiting catalytic cycles: a broader view through the energy span model. ACS Catal. 2020, 10, 12627–12635. 10.1021/acscatal.0c02332. [DOI] [Google Scholar]
  134. Su X. J.; Gao M.; Jiao L.; Liao R.-Z.; Siegbahn P. E. M.; Cheng J.-P.; Zhang M.-T. Electrocatalytic water oxidation by a dinuclear copper complex in a neutral aqueous solution. Angew. Chem., Int. Ed. 2015, 54, 4909–4914. 10.1002/anie.201411625. [DOI] [PubMed] [Google Scholar]
  135. 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. 10.1021/jacs.8b00032. [DOI] [PubMed] [Google Scholar]
  136. Shi J.; Guo Y.-H.; Xie F.; Chen Q.-F.; Zhang M.-T. Redox-active ligand assisted catalytic water oxidation by a RuIV=O intermediate. Angew. Chem., Int. Ed. 2020, 59, 4000–4008. 10.1002/anie.201910614. [DOI] [PubMed] [Google Scholar]
  137. Zhang H.-T.; Su X.-J.; Xie F.; Liao R.-Z.; Zhang M.-T. Iron-catalyzed water oxidation: O-O bond formation via intramolecular oxo-oxo interaction. Angew. Chem., Int. Ed. 2021, 60, 12467–12474. 10.1002/anie.202100060. [DOI] [PubMed] [Google Scholar]
  138. Chen Q.-F.; Xiao Y.; Liao R.-Z.; Zhang M.-T. Trinuclear nickel catalyst for water oxidation: intramolecular proton-coupled electron transfer triggered trimetallic cooperative O–O bond formation. CCS Chem. 2022, 5, 245–256. 10.31635/ccschem.022.202101668. [DOI] [Google Scholar]
  139. Zhang H.-T.; Guo Y.-H.; Xiao Y.; Du H.-Y.; Zhang M.-T. Heterobimetallic NiFe cooperative molecular water oxidation catalyst. Angew. Chem., Int. Ed. 2023, 62, e202218859 10.1002/anie.202218859. [DOI] [PubMed] [Google Scholar]
  140. Shi J.; Guo Y.-H.; Xie F.; Zhang M.-T.; Zhang H.-T. Electronic effects of redox-active ligands on ruthenium-catalyzed water oxidation. Chin. J. Catal. 2023, 52, 271–279. 10.1016/S1872-2067(23)64497-3. [DOI] [Google Scholar]
  141. Chen Q.-F.; Xian K.-L.; Zhang H.-T.; Su X.-J.; Liao R.-Z.; Zhang M.-T. Pivotal role of geometry regulation on O–O bond formation mechanism of bimetallic water oxidation catalysts. Angew. Chem., Int. Ed. 2024, 63, e202317514 10.1002/anie.202317514. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

au4c00047_si_001.pdf (19MB, pdf)

Articles from JACS Au are provided here courtesy of American Chemical Society

RESOURCES