Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

pubs.acs.

org/JACS Article

Electrocatalytic Interconversions of CO2 and Formate on a Versatile


Iron-Thiolate Platform
Yongxian Li, Jia-Yi Chen, Xinchao Zhang, Zhiqiang Peng, Qiyi Miao, Wang Chen, Fei Xie,
Rong-Zhen Liao,* Shengfa Ye,* Chen-Ho Tung, and Wenguang Wang*
Cite This: J. Am. Chem. Soc. 2023, 145, 26915−26924 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Exploring bidirectional CO2/HCO2− catalysis holds


Downloaded via INDIAN INST OF TECH KANPUR on April 1, 2024 at 11:10:06 (UTC).

significant potential in constructing integrated (photo)-


electrochemical formate fuel cells for energy storage and
applications. Herein, we report selective CO2/HCO2− electro-
chemical interconversion by exploiting the flexible coordination
modes and rich redox properties of a versatile iron-thiolate
platform, Cp*Fe(II)L (L = 1,2-Ph2PC6H4S−). Upon oxidation,
this iron complex undergoes formate binding to generate a diferric
formate complex, [(L−)2Fe(III)(μ-HCO2)Fe(III)]+, which exhibits
remarkable electrocatalytic performance for the HCO2−-to-CO2
transformation with a maximum turnover frequency (TOFmax)
∼103 s−1 and a Faraday efficiency (FE) ∼92(±4)%. Conversely, this iron system also allows for reduction at −1.85 V (vs Fc+/0) and
exhibits an impressive FE ∼93 (±3)% for the CO2-to-HCO2− conversion. Mechanism studies revealed that the HCO2−-to-CO2
electrocatalysis passes through dicationic [(L2)−•Fe(III)(μ-HCO2)Fe(III)]2+ generated by unconventional oxidation of the diferric
formate species taking place at ligand L, while the CO2-to-HCO2− reduction involves a critical intermediate of [Fe(II)-H]− that was
independently synthesized and structurally characterized.

■ INTRODUCTION
Metalloenzymes set high standards for synthetic catalysts to
“reversible catalysis” defined by Fourmond and Léger,15,16
electrocatalysts capable of performing both reduction and
perform green energy conversion by utilizing atmospheric gas oxidation reactions are considered bidirectional, but they are
molecules as sustainable feedstocks.1−3 For instance, metal- considerably reversible when they operate within a small
dependent formate dehydrogenases (FDHs) containing either departure from the equilibrium potential in both directions.17,18
molybdenum or tungsten can efficiently catalyze both formate Yang presented an exemplary reversible catalyst, [Pt(depe)2]2+,
oxidation and CO2 reduction, both having impressive reaction which functions both CO2-to-HCO2− reduction and HCO2−-to-
rates, i.e., 3400 s−1 at pH 7.5 for the former process and 280 s−1 CO2 oxidation through sharing a common Pt−H intermediate
for the latter.4,5 The microscopic CO2/HCO2− interconversions (Figure 1a).17,18 In principle, the facile occurrence of such
inspire research into coupling electricity production with CO2 catalytic processes requires that each elementary step thereof
reduction to store the energy into formic acid as a renewable should have a minimum free energy change and should not
feedstock for fuel cells (eq 1).6−8 This approach holds great entail a high barrier.16,17 Apparently, it is difficult to design
promise for constructing integrated photoelectrochemical systems to meet these requirements. As a consequence, only two
devices that harness solar energy to drive CO2 reduction and molecular bidirectional catalytic systems have been published
subsequently release electrical energy via oxidation of formate to thus far, i.e., [Pt(depe)2]2+17 and (PONOP)Ir(H)3,19 whereas
CO2 with the same device.7,9−11 Therefore, the exploration of numerous catalysts have already been reported for each
catalytic systems capable of performing electrochemical CO2/ individual transformation.12,20−25 It should be noted that for
HCO2− interconversion is of crucial importance and scientific both CO2 reduction and formate oxidation, the vast majority of
intrigue.12−14 catalysts are noble metal complexes that are typically superior to
Eox
HCO2 XoooooooooY CO2 + 2e + H+
Ered Received: September 7, 2023
(singlecatalyst) (1) Revised: November 15, 2023
Inspired by Nature’s strategy using FDHs, the straightforward Accepted: November 16, 2023
route to achieve electrocatalytic CO2/HCO2− interconversion is Published: November 29, 2023
to invoke reversible catalysis with a single metal complex.
Regarding the concepts of “bidirectional catalysis” and

© 2023 American Chemical Society https://doi.org/10.1021/jacs.3c09824


26915 J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 1. (a) Proposed catalytic cycle for reversible transformations of CO2/HCO2− by [Pt(depe)2](PF6)2. Reproduced with permission from ref 17.
Copyright 2020 John Wiley and Sons. (b) Hypothesized electrochemical CO2/HCO2− interconversions using a single catalyst. (c) Present versatile
iron-thiolate platform for the CO2/HCO2− interconversions.

Figure 2. Electrochemical analysis of (a) formate oxidation and (b) CO2 reduction by using 1-N2 as a catalyst in the THF solution. For (a), CVs of
n
Bu4NHCO2·HCO2H (gray line), 1-N2 (1 mM) (red line), and 1-N2 with nBu4NHCO2·HCO2H (blue line); for (b), CVs of 1-N2 (1 mM) under N2
(red line), under N2 with TFE (violet line), under CO2 (green line), and under CO2 with TFE (orange line). Conditions: 1 mM sample in 0.3 M
n
Bu4NPF6-THF, scan rate 0.1 V/s; glassy carbon working electrode, Ag wire reference electrode, and Pt wire counter electrode.

their 3d counterparts.20,25 More critically, the catalytic activities catalysts.20,24 The implementation of this strategy presents a
of the artificial systems in both directions are far below those of remarkable opportunity to utilize base metals for multielectron
FDHs. This reflects the challenges of reversible catalysis using a and multistep catalysis.13,27,28 The hypothesis is that the
single metal catalyst to perform CO/HCO2− conversions.2 oxidized catalyst is capable of binding formate in a bimolecular
Instead of exploring reversible catalysis, we sought to realize way that facilitates electron and proton transfer to improve rates
bidirectional conversions between CO2 and HCO2− by of the electrochemical HCO2−-to-CO2 oxidation (Figure 1b),
effectively exploiting the versatile coordination chemistry and while upon 1e− reduction, the catalyst undergoes protonation
the rich redox properties of a well-defined base metal catalyst with weak acid and can efficiently convert to metal hydride
such that the two transformations neither follow the same route species, which can be further reduced at a mild reduction
nor share the same metal hydride intermediate.7 The two potential to perform selective CO2-to-HCO2− reduction.
transformations are thus detangled and can be easily tuned by Herein, we report the first iron−sulfur system, Cp*Fe(1,2-
the substrate coordination. In this case, the catalysis could Ph2PC6H4S),29,30 which displays intriguing bidirectional elec-
proceed via stepwise electron transfer (E) to achieve chemical trocatalytic activity toward CO2/HCO2− interconversions
substrate binding and proton transfer steps (C),26 rather than (Figure 1c). Upon oxidation, the iron complex readily interacts
undergoing successive two-electron transfers prior to any with formate to generate a diferric bridging formate species,
chemical steps, as observed for the most noble metal which can be further oxidized for C−H bond cleavage. This
26916 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 3. Electrochemical analysis of formate oxidation and characterizations of bridging formate diiron complexes. (a) CVs of [1-HCO2][BAr4F] (1
mM in THF) with varying [nBu4NHCO2·HCO2H] (0−350 mM) at a scan rate of 0.1 V/s and plot of the catalytic current versus [nBu4NHCO2·
HCO2H] (inset). (b) Plot of the turnover frequency versus [nBu4NHCO2·HCO2H]. (c) CVs of [1-HCO2][BAr4F] (1 mM) with [nBu4NHCO2·
HCO2H] (350 mM) in the THF solution with a varying scan rate and the plot of the catalytic current versus the scan rate (0.1−1 V/s). Conditions for
CVs: a glassy carbon electrode was used as the working electrode, a Pt wire as the counter electrode, and a Ag wire as the reference electrode; nBu4NPF6
(0.3 M) was used as the supporting electrolyte. (d) X-ray structures of [1-HCO2]+ and [1-HCO2]2+ cations with 50% probability thermal ellipsoids.
For clarity, the BAr4F counterions, H atoms at the Cp*, and phenyl rings were omitted; Cp* and the two phenyl rings at the phosphorus site were drawn
in lines. (e) Summary of the X-ray structural and 57Fe Mössbauer parameters of [1-HCO2]+ and [1-HCO2]2+. (f) Zero-field 57Fe Mössbauer spectra of
[1-HCO2][BAr4F] and [1-HCO2][BAr4F]2 recorded at 120 K. The solid lines represent the best fit obtained with the parameters listed in Table (e).

bimetallic reaction mode represents the fastest record for carbon electrode was observed at an onset potential of 0.50 V
HCO2−-to-CO2 electrocatalysis. At a relatively low voltage, this versus Fc+/0 in the absence of the iron complex.
iron−sulfur system is capable of catalyzing highly selective CO2- Complex 1-N2 was also identified as an active electrocatalyst
to-HCO2− conversion, and the key intervening intermediate of for CO2 reduction, with CF3CH2OH (TFE, pKa = 24.0 in
an [Fe(II)-H]− monomer was independently synthesized and DMF)32,33 acting as the proton source. When a THF solution of
structurally characterized. Comprehensive experimental and 1-N2 was saturated with CO2, the reduction wave at −1.85 V
theoretical investigations offer valuable insights into such a became irreversible and slightly shifted to a less positive
bidirectional 2e−/H+ transformation on the multifunctional potential (Figure 2b). Notably, adding TFE into a CO2-
iron-thiolate platform. saturated solution resulted in a significant increase in the
cathodic current at the onset potential of −1.65 V vs Fc+/0, and
■ RESULTS AND DISCUSSION
Bidirectional CO2/Formate Electrocatalysis with 1-N2.
the reduction current rises to ca. 3-fold of that generated by [1-
N2]−/0 under CO2. In the absence of CO2, the CV of 1-N2 with
TFE showed an enhanced reduction wave shifted to a more
We initiated a study by assessing the catalytic performance of 1- negative onset potential of −1.80 V. These CV observations
N2 in formate oxidation and CO2 reduction through cyclic indicate that the iron catalyst achieves selective electrochemical
voltammetry (CV) measurements, respectively. Under a N2 CO2 reduction over proton reduction when TFE is employed as
atmosphere, the CV of 1-N2 (1 mM) in a THF-nBu4NPF6 the proton source. The preliminary results from cyclic
solution (0.3 M) displayed two reversible events at −1.85 V (ipc/ voltammetry experiments demonstrate the bidirectional electro-
ipa = 0.98) and −0.33 V (ipa/ipc = 0.95) vs Fc+/0, which are catalytic activity of the iron-thiolate complex in CO2/HCO2−
tentatively assigned to FeII/I and FeII/III redox couples (Figure 2). interconversion. These findings motivated us to investigate the
Upon addition of nBu4NHCO2·HCO2H as the formate source, electrochemical kinetics and delve into the underlying reaction
the couple at −0.33 V was eliminated, and a new redox event mechanism, which are discussed below.
emerged with a significant increase in the anodic current at an Formate Oxidation. To achieve electrocatalytic CO2/
onset potential of 0.30 V. These electrochemical observations HCO2− interconversion, the critical task lies in effectively
suggest the occurrence of a catalytic process for formate realizing the C−H bond formation and breakage through
oxidation.31 In contrast, the oxidation of formate with a glassy consecutive 2e−/H+ transfer steps with a single metal
26917 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

complex.34,35 Regarding the electrochemical formate oxidation, ij i yz


2 2
ij 0.4463 yz n p3Fv
Appel and Kubiak et al. reported the first nickel electrocatalysts TOF = kobs = jjjj cat zzzz jj
jj n
zz
zz
[Ni(P2RN2R′)2(CH3CN)]2 for HCO2−-to-CO2 oxidation (with j ip z k cat { RT
k { (3)
a maximum TOF ∼ 16 s−1), and the C−H cleavage was
proposed through binding formate to the nickel center, followed In Figure 3b, the TOF of the catalyst is plotted against
by a rate-limiting proton and two-electron transfer step to [nBu4NHCO2·HCO2H]. Initially, The TOF increases linearly
liberate CO2.31,36 Based on the preliminary analysis, we with increasing [nBu4NHCO2·HCO2H] and reaches a plateau at
hypothesized that the electrocatalytic formate oxidation is 300 mM, indicating that the TOF becomes independent of the
formate concentration. The linear region of this plot indicates a
triggered by the oxidation of 1-N2, which is followed by a
first-order dependence of the TOF on [nBu4NHCO2·HCO2H]
chemical transformation with the substrate. Typically, the
when it is below 300 mM. The dependence of the catalytic
binding of formate to metal catalysts represents a crucial step
current on the scan rate was also investigated with the
in electrocatalytic formate oxidation. [nBu4NHCO2·HCO2H] of 350 mM. As shown in Figure 3c,
To verify the proposition, we therefore attempted to the catalytic current becomes scan rate-independent when the
synthesize the corresponding iron formate complex, [1- scan rate is increased to 0.6 V/s, indicating that the catalysis is
HCO2]+. Oxidation of 1-N2 by [Cp2Fe][BAr4F] (where Cp = operating in the pure kinetic regime, where the substrate
C5H5−, BAr4F = (3,5-(CF3)2C6H3)4B) in MeCN generated the consumption is negligible.41,42 The icat/ip value, determined
corresponding ferric nitrile complex, [Cp*Fe(1,2-Ph2PC6H4S)- from the CV, was found to be 50, and a TOFmax value of 2.91 ×
(NCMe)]+ ([2-NCMe]+), in a high yield. Treating [2-NCMe]+ 103 s−1 was calculated. Furthermore, the turnover frequency was
with HCO2Na in THF caused a gradual color change from deep also determined from the preparative-scale electrolysis of
blue to purple and finally afforded the target formate complex, formate oxidation.43−45 This analysis yielded a TOFCPE value
[1-HCO2][BAr4F]. X-ray structural analysis revealed that [1- of 1.05 × 103 s−1, which aligns with the value obtained from the
HCO2][BAr4F] is a cationic diiron complex wherein the two icat/ip method. Remarkably, the present iron-based catalysis
Cp*Fe(1,2-Ph2PC6H4S) fragments are bridged by a formate exhibits the highest reaction rate for formate oxidation
ligand (Figures 3d and S60) with C−O distances of 1.255(4) compared to related homogeneous catalysts ever re-
and 1.250(4) Å, and ∠O−C−O = 121.7(3)°. In the Fourier- ported.17,19,31,36
transform infrared (FT-IR) spectrum, a weak νC−O band was Controlled potential electrolysis (CPE) was conducted on a
displayed at 1560 cm−1,37 which shifts to 1522 cm−1 for the 13C- THF solution containing 100 mM [nBu4NHCO2·HCO2H] and
labeled formate isotopologue, i.e., [1-H13CO2][BAr4F] (Figure 1 mM [1-HCO2][BAr4F]. At a potential of 0.35 V vs Fc+/0, a total
S47). charge of 12.4 C was passed through the system over the course
Complex [1-HCO2][BAr4F] in THF-nBu4NPF6 at a scan rate of the 6 h bulk electrolysis. Notably, the current remained almost
of 100 mV/s showed an irreversible redox event at 0.30 V vs constant at 0.6 mA throughout the electrolysis process (Figure
Fc+/0, tentatively attributed to a FeIII,III/III,IV redox couple, which S17). Analyses of the headspace gas with GC indicated that the
becomes increasingly reversible at higher scan rates (Figure S8). reaction produced 59.0 μmol of CO 2 (TON ∼ 12),
More important is that the anodic current rises upon adding corresponding to a Faraday efficiency of 91.8(±4)%. X-ray
n
Bu4NHCO2·HCO2H. Increasing [nBu4NHCO2·HCO2H] photoelectron spectroscopy (XPS) and rinse testing on the
leads to a gradual shift of the anodic wave to more positive glassy carbon working electrode confirmed the absence of active
potentials accompanied by the current ascending (Figure 3a). iron species attached to the electrode surface (Figures S49 and
When [nBu4NHCO2·HCO2H] reached 300 mM, the catalytic S50).46 These findings collectively demonstrated that [1-
HCO2][BAr4F] is a highly selective and robust catalyst for
current was independent of its concentration. Alternatively, at a
electrochemical formate oxidation.
constant concentration (100 mM) of nBu4NHCO2·HCO2H, the
According to the electrochemical analysis, the iron-catalyzed
anodic current also increases linearly with the concentration of
formation oxidation is based on the oxidation of [1-HCO2]-
[1-HCO2][BAr4F] over the range of 0.5−2.5 mM (Figures S15 [BAr4F]. To ensure the uncommon one-electron oxidation of
and S16). All of these results indicate the occurrence of this cationic diferric complex, bulk chemical oxidation was
electrochemical catalysis in formate oxidation.31 To analyze the subsequently carried out (eq 4). Treating a fluorobenzene
kinetics of the electrochemical formate oxidation, the catalytic solution of [1-HCO 2 ][BAr 4 F ] with one equiv of
current enhancement, icat/ip (where icat and ip represent the [(C6F5Cp)2Fe][BAr4F] (E1/2 = 0.345 V vs Fc+/0) at −30 °C
current in the presence and absence of the substrate, resulted in a rapid color change from violet to black. After being
respectively), was converted into an observed rate constant warmed to room temperature, the reaction mixture was dried
(kobs). This conversion was done using eq 2, which incorporates under vacuum. The crude product, [1-HCO2][BAr4F]2, was
Faraday’s constant (F), the universal gas constant (R), the washed with hexane and purified through crystallization from a
temperature (T), the diffusion coefficient (D), the area of the fluorobenzene/hexane solution at −30 °C.
working electrode (A), the number of electrons transferred in
the catalytic process (ncat), the number of electrons transferred [1 HCO2 ]+ e [1 HCO2 ]2 + (4)
in the redox event (np), and the scan rate (v).31,38,39
Additionally, a catalytic turnover frequency (TOF) was [1 HCO2 ]2 + + collidine CO2 (5)
calculated using eq 3 (the Supporting Information, page
X-ray crystallographic analysis revealed the solid-state
S8).31,38−40
structure of [1-HCO2][BAr4F]2, presumably a formate bridged
dicationic [Fe(III)(μ-HCO2)Fe(IV)]2+ species (Figures 3d and
icat ncat ij RT yz
jj z S61). Except for the counteranions, the framework of [1-
= jj 3 zzzkobs HCO2]2+ is identical to that of [1-HCO2]+, but the one-electron
ip 0.4463 j np Fv z
k { (2) oxidation causes slightly asymmetric binding of the formate
26918 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 4. Schematic molecular orbital diagrams and spin density plots of (a) [1-HCO2]+ and (b) [1-HCO2]2+.

ligand to the two iron centers as evidenced by the different Fe1− FeIII complex.26 These findings evidenced that [1-HCO2]+ and
O1 (1.989(3) Å) and Fe2−O2 distances (1.905(2) Å). [1-HCO2]2+ contain two low-spin ferric centers each, and their
Analogously, the two C−O bond lengths (1.255(4) and interconversion is a ligand-centered redox process.
1.250(4) Å) of the formate ligand in [1-HCO2]+ are almost The electronic structure analyses mentioned above were
identical, whereas those (1.272(4) and 1.250(4) Å) in [1- confirmed by density functional theory (DFT) calculations,
HCO2]2+ differ by ca. 0.02 Å, although the O−C−O bond angle which successfully replicated the experimentally observed
of 121.7(3)° in the former complex is comparable with that of Mössbauer parameters. Specifically, the triplet ground state of
121.5(3)° in the latter (Figure 3e). Notably, this new iron [1-HCO2]+ was found to consist of two ferromagnetically
species readily furnishes CO2 in the presence of a base (eq 5). coupled low-spin ferric centers and the broken-symmetry singlet
Treatment of [1-HCO2]2+ with one equiv of 2,4,6-trimethylpyr- state that represents the corresponding antiferromagnetic
idine (pKa = 8.1 in THF) produced CO2 in 83(±5)% yield as coupling is nearly isoenergetic with the triplet state (Figure
determined by GC analyses. This finding revealed that [1- 4a). Consequently, the computed key geometric parameters of
HCO2]2+ is indeed an intervening intermediate in the catalytic both states are almost identical, as are their estimated Mössbauer
cycle of formate oxidation. parameters.
Electronic Structure Analyses for [1-HCO2]2+ and [1- For the doublet state of [1-HCO2]2+, DFT computations
HCO2]+. To elucidate the electronic structures of [1-HCO2]+ predicted two different electronic structures; namely, one is best
and [1-HCO2]2+, we undertook comprehensive spectroscopic described as two ferromagnetically coupled low-spin ferric
characterizations. Variable temperature magnetic susceptibility centers that are antiferromagnetic coupled to a thiyl radical
(χ) measurements using a superconducting quantum interfer- delocalized onto the two supporting ligands denoted by
ence devices (SQUID) unequivocally established that [1- BS(2,1)-A, and the other as two antiferromagnetically coupled
HCO2]+ possesses an S = 1 ground state with an energetically low-spin ferric centers that interact with the same delocalized
closely spaced S = 0 excited state. Complex [1-HCO2]2+ features thiyl radical, denoted by BS(2,1)-B (Figure S63). The
an S = 1/2 ground state. Furthermore, the electron paramagnetic corresponding quartet state of [1-HCO2]2+ is distinguished by
resonance (EPR) spectrum of [1-HCO2]2+ registered a highly the three unpaired spins aligning up. These three states were
anisotropic rhombic signal with g = 2.00, 2.03, and 2.31 (Figure estimated to be essentially isoenergetic within the uncertainty
S56). Interestingly, [2-NCMe]+ elicited a similar EPR spectrum range of DFT calculations.48 In line with their energy near-
with g = 2.00, 2.03, and 2.35 (Figure S55), typical of degeneracy, the computed geometries of the three states are all
mononuclear S = 1/2 ferric complexes.47 Zero-field 57Fe in reasonable agreement with those of the experiment. However,
Mössbauer spectra of [1-HCO2]+ and [1-HCO2]2+ showed because EPR measurements disclosed that the sole unpaired
only one well-resolved quadrupole doublet (Figure 3f); electron of [1-HCO2]2+ is primarily localized on a low-spin ferric
therefore, their respective two Fe centers are more or less center, the BS(2,1)-A solution should represent its real
equivalent. More importantly, both species have nearly identical electronic structure (Figure 4b). As depicted in Figure 4b,
isomer shifts (δ = 0.43, 0.43 mm/s) and quadrupole splittings upon oxidation, the spin populations of the two Fe centers are
(ΔEQ = 1.38, 1.29 mm/s), which are comparable to those exactly the same, but those of the two S atoms substantially
measured for [Cp*FeIII(1,2-Ph2PC6H4NH)(NH3)]+ (δ = 0.40 decrease from 0.14 to −0.49, along with appreciable negative
mm/s and ΔEQ = 1.35 mm/s), a related mononuclear S = 1/2 spin densities being delocalized on the two phenyl rings
26919 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 5. Electrocatalytic CO2 reduction with 1-N2. (a) CVs of 1-N2 recorded in the CO2-saturated THF solution in the presence of varying [TFE]
(0−125 mM), (b) plots of the turnover frequency versus [TFE], and (c) plots of icat/ip versus [TFE]1/2 to determine kH/kD.

attached. The oxidation is, therefore, best interpreted as being (Figure S34). The total catalytic turnover number (TON) for
primarily ligand-centered and does not engender highly covalent formate production was calculated to be 4.9, corresponding to a
Fe(IV)-S bonding interactions. Faraday efficiency of 93(±3)%. A minor amount of H2 was also
CO2 Reduction. As mentioned above, 1-N2 also serves as an detected by GC-TCD, accounting for approximately 4.5(±1)%;
active catalyst for electrochemical CO2 reduction; the following no CO was detected (Figure S35). By contrast, when the CPE
electrochemical analyses were conducted to investigate the was conducted in the absence of 1-N2, neither formate nor H2
kinetics of CO2 reduction.13,41 As shown in Figure 5a, a addition was generated.
of TFE to the CO2-saturated solution of 1-N2 (1 mM in THF) Prior research has highlighted difficulties in achieving
resulted in a continuous increase in the catalytic current, selective production of formate over CO and H2 from
accompanied by an anodic shift in the wave potentials. The electrocatalytic CO2 reduction21,22,54,55 because the reaction
catalytic current reached a limiting value at [TFE] of 125 mM. paths heavily depend on the redox property of the catalyst and
At this stage, the catalytic cyclic voltammogram exhibited a the acidity of the proton source.23,32,43,56 For formate
characteristic S-shape response, and the catalytic current was production, the majority of reactions involve crucial metal-
also found to be independent of the scan rate (Figure S24). hydride (M-H) species generated by protonation of the reduced
These observations suggest that the catalysis is operating in the catalyst.7,12,22 However, subsequent CO2 reduction by M-H to
pure kinetic regime.41,42 To quantify the catalytic current furnish HCO2− encounters competition with proton reduction
enhancement (icat/ip) and the TOF, eqs 2 and 3 were employed, to produce H2.57 Consequently, only a limited number of
respectively. Plotting the TOF against [TFE] revealed a linear homogeneous catalysts have been reported to exhibit high
increase of the TOF with [TFE], which leveled off at [TFE] = (>90%) FE for selective CO2-to-HCO2− reduction.42,58,59
125 mM (Figure 5b). Under these conditions, TOFmax was Notably, the incorporation of a basic site into the catalysts′
determined to be 1.05 × 102 s−1, which is consistent with the coordination sphere enables the utilization of abundant metals
TOF value of 1.22 × 102 s−1 obtained from the preparative-scale such as Fe,35,49,51,53,57,60 Co,42,52 and Ni61 for the selective
electrolysis experiments (the Supporting Information, page S9 formate production at mild potentials.22,43 Remarkably, recent
and Figure S27).43−45
research on iron−porphyrin systems yielded valuable mecha-
Additionally, when the concentration of TFE is below 125
nistic and synthetic insights and established the structure−
mM, the catalytic current enhancement (icat/ip) varies linearly
property relationships that govern selectivity for CO2-to-
with the square root of the CF3CH2OH concentration
HCO2− reduction.12,53,60,62
([TFE]1/2), which indicates that the electrocatalytic process is
According to the CV analyses, the present iron-based catalysis
first-order with respect to [TFE] (Figure 5c).49,50 At a constant
[TFE] of 28 mM, we also found that the catalytic current (icat) is initiated by the reduction of 1-N2. We propose an anionic
increases linearly as a function of [Fe] in the range of 0.5−2.5 iron(II)-hydride intermediate [Cp*FeH(1,2-Ph2PC6H4S)]−
mM (Figures S25 and S26). When CF3CH2OD was used as the (abbreviated to [2-H]−) responsible for selective CO2 reduction
proton source instead of CF3CH2OH, the catalysis progressed at (for detailed information, refer to Figure 7). Particularly, the
much slower rates (Figure S21). A primary H/D kinetic isotope hypothesized anionic hydride can be accessed by employing an
effect (KIE) value of 4.8 ± 0.5 was obtained (Figure 5c). A KIE alternative method to incorporate a hydride ligand onto the
value of such a magnitude is close to the reported KIE values Cp*FeL platform (eq 6). The addition of LiBHEt3 (1 M in
observed in electrochemical CO2-to-HCOOH reduction THF) to a THF solution of 1-N2 at −30 °C resulted in a rapid
catalyzed by [CpCo(P2BuN2Cy)I]+ (5.0 ± 0.4)42 and Fe- transformation of the brown solution to deep green. Following
(tbudhbpy)Cl (4.8 ± 0.9),51 both of which involve a metal the workup, the desired hydride product, [2-H]−, was obtained
hydride (M-H) intermediate.20,23,52,53 by crystallization from the ether solution in 70 (±5)% yield. The
To determine the product of the electrocatalytic CO2
1
H NMR spectrum of [2-H]− displays a characteristic hydride
reduction, a CPE experiment was carried out at −1.85 V (vs doublet at δ − 18.19 with JH,P = 64.2 Hz, while the phosphorus
Fc+/0) in a CO2-saturated THF-nBu4NPF6 (0.3 M) solution resonance is observed at δ 106 in 31P NMR. The solid-state
containing 1 mM 1-N2 and 50 mM TFE. Over the course of 4 h structure of [2-H]Li, determined by X-ray crystallographic
electrolysis, a total charge of 10.2 C was passed through the analysis, reveals an anion−cation framework composed of
system. After the reaction, the product was analyzed by 1H NMR [Cp*FeH(1,2-Ph2PC6H4S)]Li(THF)(Et2O) (Figure 6). The
spectroscopy, which confirmed the production of formate position of the hydride ligand was located by inspection of the
26920 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

difference density map, and the Fe−H bond length was refined
to be 1.46(3) Å.
1 N2 + LiBHEt 3 [2 H]Li (6)

[2 H]Li + CO2 2 + HCO2 Li (7)

As anticipated, [2-H]Li reacts with CO2 to produce formate


(eq 7). When a THF solution of [2-H]Li was exposed to CO2, its
color immediately changed from dark green to brown. The
reaction yielded formate salt in 72% yield, as determined by 1H
NMR analyses (Figure S36). The organoiron byproduct,
presumably Cp*Fe(1,2-Ph2PC6H4S) (2), was further trapped
by CO to furnish a stable adduct, Cp*Fe(1,2-Ph2PC6H4S)(CO)
Figure 6. X-ray structure of [2-H]Li with 50% probability of thermal (Figure S37).28 These results confirm that [Cp*FeH(1,2-
ellipsoids. Ph2PC6H4S)]− is highly hydridic and capable of triggering the
conversion CO2 to formate (ΔGH− = 44 kcal/mol in MeCN).55

Figure 7. Proposed mechanisms and the transition states (TS1 and TS2) for electrocatalytic formate oxidation and CO2 reduction with the iron-
thiolate complex. All reaction and activation energies are given in units of kcal/mol, and the experimental redox potentials are referenced to the Fc+/0
couple.

26921 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Mechanism. Based on the experimental observations, formate complex that exhibits efficient electrochemical catalytic
computational studies were conducted to gain further activity with a maximum turnover frequency of ∼103 s−1.
mechanistic insights into the iron-based bidirectional CO2/ Remarkably, the resulting diiron system undergoes further one-
formate electrocatalysis (Figure 7). Regarding the formate electron oxidation, leading to an uncommon [(L2)−•Fe(III)(μ-
oxidation, starting from [1-HCO2]2+, the second HCO2− HCO2)Fe(III)]2+ dicationic radical species, which allows for
molecule functions as a Brønsted base and deprotonates the facile deprotonation of the HCO2− ligand under mild conditions
bridging HCO2− ligand, which is coupled with a single electron to liberate CO2. Upon reduction, the iron-thiolate monomer
transfer (SET) from the formate substrate to the supporting exhibits ∼93(±3)% FE in electrocatalytic CO2 reduction to
ligand radical cation. This process has to overcome a barrier formate. The reaction likely passes through an anionic [Fe(II)-
(TS1) of +6.8 kcal/mol in the doublet relative to [1-HCO2]2+ H]− intermediate, with the ensuing hydride transfer represent-
plus HCO2 − and results in the formation of doublet ing the rate-limiting step, as evidenced by a primary H/D KIE
intermediate Int1. For 2TS1 (Figure S64), the spin populations value of 4.8(±0.5). Combining electrochemical kinetic studies
on Fe1, Fe2, and the (HCO2)2 moieties are 1.13, −1.09, and and stoichiometric reaction sequences with DFT studies, the
0.85, respectively. Hence, the electronic structure of 2TS1 can be mechanistic intricacies for the 2e−/H+ bidirectional CO2-
described as the (HCO2−)2 radical anion ferromagnetic formate catalysis were meticulously elucidated. In terms of
coupling with two Fe(III) centers [FeIII-(HCO2−)2•-FeIII]+. As carbon efficiency, however, the current electrocatalytic HCO2-
shown in Figure 7, the distances of the C1−H1 and H1−O3 to-CO2 oxidation system necessitates the addition of an extra 1
bonds are 1.26 and 1.33 Å in 2TS1. The formed CO2 links the equiv of formate as a Lewis base. This, to some extent, reduces
iron centers, and HCO2H interacts with the S atom of the the reaction efficiency, highlighting a potential area for
catalyst through the hydrogen bonding interaction. The spin improvement in future systems.
populations on Fe1, Fe2, and CO2 are 1.15, −1.15, and 0.91,
respectively (Figure S64). Based on the spin population analysis
of Int1, the CO2 moiety (total spin population of 0.91) is

*
ASSOCIATED CONTENT
sı Supporting Information
suggested to be a radical anion. Finally, a SET from CO2•− to The Supporting Information is available free of charge at
Fe1 takes place and releases CO2 and HCO2H, accompanied by https://pubs.acs.org/doi/10.1021/jacs.3c09824.
the generation of ferrous and ferric monomer catalysts 2 and
[2]+. The process was calculated to be exergonic by 36.1 kcal/ Discussions of experimental details, synthesis and
mol. This formate oxidation pathway is reminiscent of the characterization information, electrochemical studies,
orthogonal hydride (electron/proton) transfer mechanism for spectroscopic data, crystal data, structure refinement
Mo-dependent formate dehydrogenase, in which the formate parameters, and DFT calculations (PDF)
proton is delivered to a Brønsted base (anionic selenium Accession Codes
cysteine residue) and the two electrons to the metal center.2 CCDC 2278772−2278773, 2278777, and 2278818 contain the
The pathway for the iron-catalyzed electrochemical reduction supplementary crystallographic data for this paper. These data
of CO2-to-formate was also elucidated by DFT calculations. Our can be obtained free of charge via www.ccdc.cam.ac.uk/
previous studies29,30 indicated that dissolution of 1-N2 in THF data_request/cif, or by emailing data_request@ccdc.cam.ac.
causes N2 expulsion and forms monomeric complex Cp*Fe(1,2- uk, or by contacting The Cambridge Crystallographic Data
Ph2PC6H4S) (2). Reduction of 2 at −1.85 V vs Fc+/0 generates Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44
catalytically active species [2]−, which allows for protonation by 1223 336033.
(CF3CH2OH)2 through transition state TS2 (Figure 7). The
barrier of TS2 is calculated to be +8.2 kcal/mol (Figure S65).
The geometric features of TS2 (Figure S66), having Fe−H1 and
H1−O1 bond distances of 1.61 and 1.46 Å, respectively, affirm
■ AUTHOR INFORMATION
Corresponding Authors
that it is on the way of a proton transfer process. This reaction Rong-Zhen Liao − School of Chemistry and Chemical
generates a neutral ferric hydride compound, 2-H, after releasing Engineering, Huazhong University of Science and Technology,
the (CF3CH2O)−HOCH2CF3. Further, one-electron reduction Wuhan 430074, China; orcid.org/0000-0002-8989-
of 2-H produces anionic hydride [2-H] − , which was 6928; Email: rongzhen@hust.edu.cn.
independently synthesized and characterized, as revealed Shengfa Ye − State Key Laboratory of Catalysis, Dalian
above. The facile insertion of CO2 into the Fe−H bond in [2- Institute of Chemical Physics, Chinese Academy of Sciences,
H]− was achieved via a barrierless diffusion-control way (Figure Dalian 116023, China; orcid.org/0000-0001-9747-1412;
S67). The formation of the formate complex [2-OCHO]− lies at Email: shengfa.ye@dicp.ac.cn.
−21.0 kcal/mol relative to [2-H]− (Figure 7). The final step of Wenguang Wang − College of Chemistry, Beijing Normal
HCO2− release from [2-OCHO]− to regenerate 2 is a slightly University, Beijing 100875, China; orcid.org/0000-0002-
endergonic process with a Gibbs free energy change (ΔG0) of 4108-7865; Email: wwg@bnu.edu.cn.
+1.5 kcal/mol. The possibility that CO2 direct attacks 2-H to Authors
generate HCO2− (via TS4, Figures S65−S66) can be ruled out Yongxian Li − College of Chemistry, Beijing Normal University,
due to the high barrier of +28.6 kcal/mol. Beijing 100875, China

■ CONCLUSIONS
We have demonstrated the first iron system that achieves
Jia-Yi Chen − School of Chemistry and Chemical Engineering,
Huazhong University of Science and Technology, Wuhan
430074, China
electrocatalytic CO2 and formate interconversion by exploiting Xinchao Zhang − State Key Laboratory of Catalysis, Dalian
the coordination chemistry and redox chemistry of a Cp*Fe(1,2- Institute of Chemical Physics, Chinese Academy of Sciences,
Ph2PC6H4S) platform. Upon oxidation, the iron-thiolate Dalian 116023, China; University of Chinese Academy of
complex readily binds formate to form a diferric bridging Sciences, Beijing 100049, China
26922 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

Zhiqiang Peng − College of Chemistry, Beijing Normal Electrocatalysis: Insights into Materials Design. Science 2017, 355,
University, Beijing 100875, China No. eaad4998.
Qiyi Miao − State Key Laboratory of Catalysis, Dalian Institute (9) Vo, T.; Purohit, K.; Nguyen, C.; Biggs, B.; Mayoral, S.; Haan, J. L.
of Chemical Physics, Chinese Academy of Sciences, Dalian Formate: An Energy Storage and Transport Bridge between Carbon
Dioxide and a Formate Fuel Cell in a Single Device. ChemSusChem
116023, China; University of Chinese Academy of Sciences,
2015, 8, 3853−3858.
Beijing 100049, China (10) Nocera, D. G. Solar Fuels and Solar Chemicals Industry. Acc.
Wang Chen − State Key Laboratory of Catalysis, Dalian Chem. Res. 2017, 50, 616−619.
Institute of Chemical Physics, Chinese Academy of Sciences, (11) Cruz, E. A. R.; Nishiori, D.; Wadsworth, B. L.; Nguyen, N. P.;
Dalian 116023, China; University of Chinese Academy of Hensleigh, L. K.; Khusnutdinova, D.; Beiler, A. M.; Moore, G. F.
Sciences, Beijing 100049, China Molecular-Modified Photocathodes for Applications in Artificial
Fei Xie − State Key Laboratory of Catalysis, Dalian Institute of Photosynthesis and Solar-to-Fuel Technologies. Chem. Rev. 2022,
Chemical Physics, Chinese Academy of Sciences, Dalian 122, 16051−16109.
116023, China (12) Saha, P.; Amanullah, S.; Dey, A. Selectivity in Electrochemical
Chen-Ho Tung − School of Chemistry and Chemical CO2 Reduction. Acc. Chem. Res. 2022, 55, 134−144.
Engineering, Shandong University, Jinan 250100, China; (13) Dubois, M. R.; Dubois, D. L. Development of Molecular
Electrocatalysts for CO2 Reduction and H2 Production/Oxidation. Acc.
orcid.org/0000-0001-9999-9755
Chem. Res. 2009, 42, 1974−1982.
Complete contact information is available at: (14) Barlow, J. M.; Yang, J. Y. Thermodynamic Considerations for
https://pubs.acs.org/10.1021/jacs.3c09824 Optimizing Selective CO2 Reduction by Molecular Catalysts. ACS Cent.
Sci. 2019, 5, 580−588.
Notes (15) Fourmond, V.; Plumeré, N.; Léger, C. Reversible Catalysis. Nat.
The authors declare no competing financial interest. Rev. Chem. 2021, 5, 348−360.
(16) Fourmond, V.; Wiedner, E. S.; Shaw, W. J.; Léger, C.

■ ACKNOWLEDGMENTS
W.W. thanks the financial support from the National Natural
Understanding and Design of Bidirectional and Reversible Catalysts
of Multielectron, Multistep Reactions. J. Am. Chem. Soc. 2019, 141,
11269−11285.
Science Foundation of China (22022102 and 22071010) and (17) Cunningham, D. W.; Barlow, J. M.; Velazquez, R. S.; Yang, J. Y.
the Natural Science Foundation of Shandong Province Reversible and Selective CO2 to HCO2− Electrocatalysis near the
(ZR2019ZD45). S.Y. gratefully acknowledges the financial Thermodynamic Potential. Angew. Chem., Int. Ed. 2020, 59, 4443−
support from the National Natural Science Foundation of 4447.
(18) Cunningham, D. W.; Yang, J. Y. Kinetic and Mechanistic Analysis
China (92161204). R.-Z.L. acknowledges the financial support of a Synthetic Reversible CO2/HCO2− Electrocatalyst. Chem. Commun.
from the National Natural Science Foundation of China 2020, 56, 12965−12968.
(21873031). (19) Bi, J.; Hou, P.; Kang, P. Single Iridium Pincer Complex for

■ REFERENCES
(1) Stripp, S. T.; Duffus, B. R.; Fourmond, V.; Léger, C.; Leimkühler,
Roundtrip Electrochemical Conversion between Carbon Dioxide and
Formate. ChemCatChem 2019, 11, 2069−2072.
(20) Francke, R.; Schille, B.; Roemelt, M. Homogeneously Catalyzed
S.; Hirota, S.; Hu, Y.; Jasniewski, A.; Ogata, H.; Ribbe, M. W. Second Electroreduction of Carbon Dioxide�Methods, Mechanisms, and
and Outer Coordination Sphere Effects in Nitrogenase, Hydrogenase, Catalysts. Chem. Rev. 2018, 118, 4631−4701.
Formate Dehydrogenase, and CO Dehydrogenase. Chem. Rev. 2022, (21) Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A Review of Catalysts for the
122, 11900−11973. Electroreduction of Carbon Dioxide to Produce Low-Carbon Fuels.
(2) Yang, J. Y.; Kerr, T. A.; Wang, X. S.; Barlow, J. M. Reducing CO2 to Chem. Soc. Rev. 2014, 43, 631−675.
HCO2− at Mild Potentials: Lessons from Formate Dehydrogenase. J. (22) Amanullah, S.; Saha, P.; Nayek, A.; Ahmed, M. E.; Dey, A.
Am. Chem. Soc. 2020, 142, 19438−19445. Biochemical and Artificial Pathways for the Reduction of Carbon
(3) Appel, A. M.; Bercaw, J. E.; Bocarsly, A. B.; Dobbek, H.; DuBois, Dioxide, Nitrite and the Competing Proton Reduction: Effect of 2nd
D. L.; Dupuis, M.; Ferry, J. G.; Fujita, E.; Hille, R.; Kenis, P. J. A.; Sphere Interactions in Catalysis. Chem. Soc. Rev. 2021, 50, 3755−3823.
Kerfeld, C. A.; Morris, R. H.; Peden, C. H. F.; Portis, A. R.; Ragsdale, S. (23) Kinzel, N. W.; Werlé, C.; Leitner, W. Transition Metal
W.; Rauchfuss, T. B.; Reek, J. N. H.; Seefeldt, L. C.; Thauer, R. K.; Complexes as Catalysts for the Electroconversion of CO2: An
Waldrop, G. L. Frontiers, Opportunities, and Challenges in Organometallic Perspective. Angew. Chem., Int. Ed. 2021, 60, 11628−
Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 11686.
2013, 113, 6621−6658. (24) Kang, P.; Chen, Z.; Brookhart, M.; Meyer, T. J. Electrocatalytic
(4) Robinson, W. E.; Bassegoda, A.; Reisner, R.; Hirst, J. Oxidation- Reduction of Carbon Dioxide: Let the Molecules Do the Work. Top.
State-Dependent Binding Properties of the Active Site in a Mo- Catal. 2015, 58, 30−45.
Containing Formate Dehydrogenase. J. Am. Chem. Soc. 2017, 139, (25) Kinzel, N. W.; Demirbas, D.; Bill, E.; Weyhermüller, T.; Werlé,
9927−9936. C.; Kaeffer, N.; Leitner, W. Systematic Variation of 3d Metal Centers in
(5) Bassegoda, A.; Madden, C.; Wakerley, D. W.; Reisner, E.; Hirst, J. a Redox-Innocent Ligand Environment: Structures, Electrochemical
Reversible Interconversion of CO2 and Formate by a Molybdenum- Properties, and Carbon Dioxide Activation. Inorg. Chem. 2021, 60,
Containing Formate Dehydrogenase. J. Am. Chem. Soc. 2014, 136, 19062−19078.
15473−15476. (26) Li, Y.; Chen, J.-Y.; Miao, Q.; Yu, X.; Feng, L.; Liao, R.-Z.; Ye, S.;
(6) Singh, A. K.; Singh, S.; Kumar, A. Hydrogen Energy Future with Tung, C.-H.; Wang, W. A Parent Iron Amido Complex in Catalysis of
Formic Acid: A Renewable Chemical Hydrogen Storage System. Catal. Ammonia Oxidation. J. Am. Chem. Soc. 2022, 144, 4365−4375.
Sci. Technol. 2016, 6, 12−40. (27) Costentin, C.; Savéant, J.-M. Multielectron, Multistep Molecular
(7) Waldie, K. M.; Brunner, F. M.; Kubiak, C. P. Transition Metal Catalysis of Electrochemical Reactions: Benchmarking of Homoge-
Hydride Catalysts for Sustainable Interconversion of CO2 and neous Catalysts. ChemElectroChem 2014, 1, 1226−1236.
Formate: Thermodynamic and Mechanistic Considerations. ACS (28) Rountree, E. S.; Dempsey, J. L. Potential-Dependent Electro-
Sustainable Chem. Eng. 2018, 6, 6841−6848. catalytic Pathways: Controlling Reactivity with pKa for Mechanistic
(8) Seh, Z. W.; Kibsgaard, J.; Dickens, C. F.; Chorkendorff, I.; Investigation of a Nickel-Based Hydrogen Evolution Catalyst. J. Am.
Nørskov, J. K.; Jaramillo, T. F. Combining Theory and Experiment in Chem. Soc. 2015, 137, 13371−13380.

26923 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924
Journal of the American Chemical Society pubs.acs.org/JACS Article

(29) Zhang, F.; Song, H.; Zhuang, X.; Tung, C.-H.; Wang, W. Iron- and Their Corresponding Heme Proteins. Coord. Chem. Rev. 1999,
Catalyzed 1,2-Selective Hydroboration of N-Heteroarenes. J. Am. 185−186, 471−534.
Chem. Soc. 2017, 139, 17775−17778. (48) Ye, S.; Neese, F. Accurate Modeling of Spin-State Energetics in
(30) Song, H.; Ye, K.; Geng, P.; Han, X.; Liao, R.; Tung, C.-H.; Wang, Spin-Crossover Systems with Modern Density Functional Theory.
W. Activation of Epoxides by a Cooperative Iron−Thiolate Catalyst: Inorg. Chem. 2010, 49, 772−774.
Intermediacy of Ferrous Alkoxides in Catalytic Hydroboration. ACS (49) Rail, D.; Berben, L. A. Directing the Reactivity of [HFe4N-
Catal. 2017, 7, 7709−7717. (CO)12]− toward H+ or CO2 Reduction by Understanding the
(31) Galan, B. R.; Schöffel, J.; Linehan, J. C.; Seu, C.; Appel, A. M.; Electrocatalytic Mechanism. J. Am. Chem. Soc. 2011, 133, 18577−
Roberts, J. A. S.; Helm, M. L.; Kilgore, U. J.; Yang, J. Y.; DuBois, D. L.; 18579.
Kubiak, C. P. Electrocatalytic Oxidation of Formate by [Ni- (50) Savéant, J.-M.; Vianello, E. Potential-Sweep Chronoamperom-
(PR2NR′2)2(CH3CN)]2+ Complexes. J. Am. Chem. Soc. 2011, 133, etry Theory of Kinetic Currents in the Case of a First Order Chemical
12767−12779. Reaction Preceding the Electron-Transfer Process. Electrochim. Acta
(32) Costentin, C.; Drouet, S.; Passard, G.; Robert, M.; Savéant, J.-M. 1963, 8, 905−923.
Proton-Coupled Electron Transfer Cleavage of Heavy-Atom Bonds in (51) Nichols, A. W.; Chatterjee, S.; Sabat, M.; Machan, C. W.
Electrocatalytic Processes. Cleavage of a C−O Bond in the Catalyzed Electrocatalytic Reduction of CO2 to Formate by an Iron Schiff Base
Electrochemical Reduction of CO2. J. Am. Chem. Soc. 2013, 135, 9023− Complex. Inorg. Chem. 2018, 57, 2111−2121.
9031. (52) Dey, S.; Todorova, T. K.; Fontecave, M.; Mougel, V.
(33) Gotico, P.; Boitrel, B.; Guillot, R.; Sircoglou, M.; Quaranta, A.; Electroreduction of CO2 to Formate with Low Overpotential Using
Halime, Z.; Leibl, W.; Aukauloo, A. Second-Sphere Biomimetic Cobalt Pyridine Thiolate Complexes. Angew. Chem., Int. Ed. 2020, 59,
Multipoint Hydrogen-Bonding Patterns to Boost CO2 Reduction of 15726−15733.
Iron Porphyrins. Angew. Chem., Int. Ed. 2019, 58, 4504−4509. (53) Margarit, C. G.; Asimow, N. G.; Costentin, C.; Nocera, D. G.
(34) Ceballos, B. M.; Yang, J. Y. Directing the Reactivity of Metal Tertiary Amine-Assisted Electroreduction of Carbon Dioxide to
Hydrides for Selective CO2 Reduction. Proc. Natl. Acad. Sci. U.S.A. Formate Catalyzed by Iron Tetraphenylporphyrin. ACS Energy Lett.
2018, 115, 12686−12691. 2020, 5, 72−78.
(35) Loewen, N. D.; Neelakantan, T. V.; Berben, L. A. Renewable (54) Johnson, B. A.; Maji, S.; Agarwala, H.; White, T. A.; Mijangos, E.;
Formate from C−H Bond Formation with CO2: Using Iron Carbonyl Ott, S. Activating a Low Overpotential CO2 Reduction Mechanism by a
Clusters as Electrocatalysts. Acc. Chem. Res. 2017, 50, 2362−2370. Strategic Ligand Modification on a Ruthenium Polypyridyl Catalyst.
(36) Seu, C. S.; Appel, A. M.; Doud, M. D.; Dubois, D. L.; Kubiak, C. Angew. Chem., Int. Ed. 2016, 55, 1825−1829.
P. Formate Oxidation via β-Deprotonation in [Ni- (55) Machan, C. W.; Sampson, M. D.; Kubiak, C. P. A Molecular
(PR2NR′2)2(CH3CN)]2+ Complexes. Energy Environ. Sci. 2012, 5, Ruthenium Electrocatalyst for the Reduction of Carbon Dioxide to CO
6480−6490. and Formate. J. Am. Chem. Soc. 2015, 137, 8564−8571.
(37) Creutz, S. E.; Peters, J. C. Diiron Bridged-Thiolate Complexes (56) Chen, L.; Guo, Z.; Wei, X.-G.; Gallenkamp, C.; Bonin, J.;
that Bind N2 at the FeIIFeII, FeIIFeI, and FeIFeI Redox States. J. Am. Anxolabéhère-Mallart, E.; Lau, K.-C.; Lau, T.-C.; Robert, M. Molecular
Chem. Soc. 2015, 137, 7310−7313. Catalysis of the Electrochemical and Photochemical Reduction of CO2
(38) Rountree, E. S.; McCarthy, B. D.; Eisenhart, T. T.; Dempsey, J. L. with Earth-Abundant Metal Complexes. Selective Production of CO vs
Evaluation of Homogeneous Electrocatalysts by Cyclic Voltammetry. HCOOH by Switching of the Metal Center. J. Am. Chem. Soc. 2015,
Inorg. Chem. 2014, 53, 9983−10002. 137, 10918−10921.
(39) Stewart, M. P.; Ho, M.-H.; Wiese, S.; Lindstrom, M. L.; (57) Taheri, A.; Carr, C. R.; Berben, L. A. Electrochemical Methods
Thogerson, C. E.; Raugei, S.; Bullock, R. M.; Helm, M. L. High for Assessing Kinetic Factors in the Reduction of CO2 to Formate:
Catalytic Rates for Hydrogen Production Using Nickel Electrocatalysts Implications for Improving Electrocatalyst Design. ACS Catal. 2018, 8,
with Seven-Membered Cyclic Diphosphine Ligands Containing One 5787−5793.
Pendant Amine. J. Am. Chem. Soc. 2013, 135, 6033−6046. (58) Kang, P.; Cheng, C.; Chen, Z.; Schauer, C. K.; Meyer, T. J.;
(40) Lee, K. J.; Elgrishi, N.; Kandemir, B.; Dempsey, J. L. Brookhart, M. Selective Electrocatalytic Reduction of CO2 to Formate
Electrochemical and Spectroscopic Methods for Evaluating Molecular by Water-Stable Iridium Dihydride Pincer Complexes. J. Am. Chem. Soc.
Electrocatalysts. Nat. Rev. Chem. 2017, 1, No. 0039. 2012, 134, 5500−5503.
(41) Savéant, J.-M. Molecular Catalysis of Electrochemical Reactions. (59) Taheri, A.; Thompson, E. J.; Fettinger, J. C.; Berben, L. A. An
Mechanistic Aspects. Chem. Rev. 2008, 108, 2348−2378. Iron Electrocatalyst for Selective Reduction of CO2 to Formate in
(42) Roy, S.; Sharma, B.; Pécaut, J.; Simon, P.; Fontecave, M.; Tran, P. Water: Including Thermochemical Insights. ACS Catal. 2015, 5, 7140−
D.; Derat, E.; Artero, V. Molecular Cobalt Complexes with Pendant 7151.
Amines for Selective Electrocatalytic Reduction of Carbon Dioxide to (60) Amanullah, S.; Saha, P.; Dey, A. Activating the Fe(I) State of Iron
Formic Acid. J. Am. Chem. Soc. 2017, 139, 3685−3696. Porphyrinoid with Second-Sphere Proton Transfer Residues for
(43) Costentin, C.; Drouet, S.; Bobert, M.; Savéant, J.-M. A Local Selective Reduction of CO2 to HCOOH via Fe(III/II)−COOH
Intermediate(s). J. Am. Chem. Soc. 2021, 143, 13579−13592.
Proton Source Enhances CO2 Electroreduction to CO by a Molecular
(61) Fogeron, T.; Todorova, T. K.; Porcher, J.-P.; Gomez-Mingot, M.;
Fe Catalyst. Science 2012, 338, 90−94.
Chamoreau, L.-M.; Mellot-Draznieks, C.; Li, Y.; Fontecave, M. A
(44) Costentin, C.; Drouet, S.; Robert, M.; Savéant, J.-M. Turnover
Bioinspired Nickel(bis-dithiolene) Complex as a Homogeneous
Numbers, Turnover Frequencies, and Overpotential in Molecular
Catalyst for Carbon Dioxide Electroreduction. ACS Catal. 2018, 8,
Catalysis of Electrochemical Reactions. Cyclic Voltammetry and
2030−2038.
Preparative-Scale Electrolysis. J. Am. Chem. Soc. 2012, 134, 11235−
(62) Costentin, C.; Passard, G.; Robert, M.; Savéant, J.-M. Pendant
11242.
Acid−Base Groups in Molecular Catalysts: H-Bond Promoters or
(45) Gotico, P.; Roupnel, L.; Guillot, R.; Sircoglou, M.; Leibl, W.;
Proton Relays? Mechanisms of the Conversion of CO2 to CO by
Halime, Z.; Aukauloo, A. Atropisomeric Hydrogen Bonding Control
Electrogenerated Iron(0)Porphyrins Bearing Prepositioned Phenol
for CO2 Binding and Enhancement of Electrocatalytic Reduction at
Functionalities. J. Am. Chem. Soc. 2014, 136, 11821−11829.
Iron Porphyrins. Angew. Chem., Int. Ed. 2020, 59, 22451−22455.
(46) Lee, K. L.; 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.
(47) Walker, F. A. Magnetic Spectroscopic (EPR, ESEEM,
Mössbauer, MCD and NMR) Studies of Low-Spin Ferriheme Centers

26924 https://doi.org/10.1021/jacs.3c09824
J. Am. Chem. Soc. 2023, 145, 26915−26924

You might also like