Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Nov 16;12(23):14689–14697. doi: 10.1021/acscatal.2c03297

Potential- and Buffer-Dependent Selectivity for the Conversion of CO2 to CO by a Cobalt Porphyrin-Peptide Electrocatalyst in Water

Jose L Alvarez-Hernandez 1, Alison A Salamatian 1, Ji Won Han 1, Kara L Bren 1,*
PMCID: PMC9724230  PMID: 36504916

Abstract

graphic file with name cs2c03297_0009.jpg

A semisynthetic electrocatalyst for carbon dioxide reduction to carbon monoxide in water is reported. Cobalt microperoxidase-11 (CoMP11-Ac) is shown to reduce CO2 to CO with a turnover number of up to 32,000 and a selectivity of up to 88:5 CO:H2. Higher selectivity for CO production is favored by a less cathodic applied potential and use of a higher pKa buffer. A mechanistic hypothesis is presented in which avoiding the formation and protonation of a formal Co(I) species favors CO production. These results demonstrate how tuning reaction conditions impact reactivity toward CO2 reduction for a biocatalyst previously developed for H2 production.

Keywords: biocatalyst, carbon dioxide reduction, cobalt porphyrin, electrocatalysis, overpotential

Introduction

Carbon dioxide (CO2) is an abundant and attractive feedstock for renewable fuels. Advances in catalysis are crucial for the development of systems for CO2 utilization,1,2 therefore attracting significant interest in the chemistry community.26 The inertness and stability of CO2 present both kinetic and thermodynamic barriers to its activation.7 The reduction of CO2 to any stable product is a multi-proton, multi-electron process (for example, see eq 1) with high activation energy, requiring effective catalysts to drive the process at acceptable rates.8 Molecular catalysts have proven successful in CO2 reduction reactions (CO2RR) and have enabled detailed mechanistic study,5,816 providing insight into the roles of both Brønsted1720 and Lewis acids,21,22 as well as the coordination of electron transfer with both proton transfer and bond breaking and formation.23 Water is a desirable solvent to use for CO2 reduction,5,2430 yet developing and understanding CO2RR catalysis in water brings several challenges: poor solubility of CO2 ([CO2] = 0.0383 M at 20 °C and 1 atm of CO2),31 pH-dependent equilibria among CO2 and its hydration products (H2CO3, HCO3, and CO32–), and competition from the hydrogen evolution reaction (HER; eq 2).

graphic file with name cs2c03297_m001.jpg 1
graphic file with name cs2c03297_m002.jpg 2

In electrocatalysis, the amount of energy beyond the thermodynamic requirements needed to drive a reaction at a given rate is described by the overpotential. Typically, lowering the overpotential for a given catalyst comes at the expense of slowing catalysis, with the log of the rate exhibiting a linear dependence on overpotential.3234 The relationship between overpotential and catalyst selectivity is a less explored topic. Studies of potential-dependent product selectivity are reported for solid (nonmolecular) electrocatalysts,3537 but studies of this nature for molecular catalysts are less common.3840

This study reports on the effect of applied potential on selectivity for CO vs H2 production from CO2 in water by a biomolecular catalyst. We describe a cobalt porphyrin with a covalently attached peptide (CoMP11-Ac; Figure 1), previously described as a catalyst for HER,4143 as an active and selective CO2-reduction electrocatalyst in water. CoMP11-Ac reaches a turnover number (TON) > 12,000 (at 2 h) for CO2 reduction to CO at an applied potential of −1.2 V (all potentials here are reported vs Ag/AgCl/KCl(1M)) with 85% faradaic yield. Our report is notable as a rare demonstration of the use of applied potential to control product selectivity for CO2RR by a molecular catalyst. Furthermore, a mechanistic proposal is put forward with the support of observed effects of potential, buffer acid pKa, and CO2 partial pressure on catalysis.

Figure 1.

Figure 1

CoMP11-Ac. Reproduced with permission from ref (43). Copyright 2020, American Chemical Society.

Results and Discussion

Effects of Potential

The CO2RR activity of CoMP11-Ac in 50 mM NaHCO3 (pKa1 6.4) solution saturated with CO2 was evaluated by cyclic voltammetry (CV) using a hanging mercury drop electrode, as shown in Figure 2. Dip-and-stir experiments44 reveal that CoMP11-Ac adsorbs to the mercury electrode, indicating that it behaves as an immobilized molecular catalyst (details in Figures S1–S4). A precedent for a system of this nature is that of Ni-cyclam, a CO2 reduction catalyst also adsorbed onto a mercury electrode.45,46 Importantly, the activity of particulate cobalt in this reaction is prevented by the use of a mercury electrode, as mercury amalgamates cobalt.4749

Figure 2.

Figure 2

CVs of 1 μM CoMP11-Ac in 50 mM NaHCO3, 0.1 M KCl, pH 6.1 ± 0.1 at 100 mV/s, under 1 atm of the indicated gas. Arrows in the CV traces indicate the scanning direction.

The CVs of CoMP11-Ac, scanning from 0 to −1.6 V (and the opposite in the return scan), show no measurable current enhancement until −1.2 V when a broad first wave is observed, followed by a second feature of higher current with a half-wave potential (Eh) of ∼ −1.4 V. The catalytic CV current for both features is significantly higher for CO2-saturated vs N2-saturated solutions at the same pH, suggesting that CoMP11-Ac may catalyze CO2RR. An inverted peak is also observed upon switching the scanning direction, which indicates that the catalyst is partially deactivated at low potentials and reactivated upon scanning anodically.50 This phenomenon has been described for other molecular catalysts based on cobalt,5153 as well as other transition metals.5456

To further characterize the activity of this cobalt porphyrin-peptide toward CO2RR, we performed controlled potential electrolysis (CPE) experiments at both −1.2 and −1.4 V in 0.5 M NaHCO3 for 2 h in solutions purged with either CO2 or N2 and under 1 atm of the purging gas (Tables 1 and S1, and Figure 3). Under a CO2 atmosphere, the charge passed with CoMP11-Ac present is comparable at both applied potentials (Figure 3), yet the product distribution is rather different. The faradic efficiency for H2 (FE(H2)) decreases from 23% at −1.4 V to 4% at −1.2 V, while FE(CO) increases from 61% at −1.4 V to 83% at −1.2 V. Furthermore, the turnover number (TON) for CO production measured after 2 h of CPE increases from 2500 at −1.4 V to 3300 at −1.2 V. In N2-saturated NaHCO3 solution at −1.4 V, CoMP11-Ac produces H2 with a 76% FE and CO with a 16% FE; CO arises from the reduction of the CO2 in equilibrium with NaHCO3 buffer.

Table 1. Summary of 2-h CPE Results for 1 μM CoMP11-Ac in 0.5 M NaHCO3a.

  E (V)b FE(H2) (%) FE(CO) (%) TON(H2) TON(CO) QT (C)
CO2 –1.4 23 ± 2 61 ± 4 1000 ± 200 2600 ± 400 4.0 ± 0.7
–1.2 4 ± 1 84 ± 13 140 ± 40 3300 ± 100 3.7 ± 0.9
N2 –1.4 77 ± 5 15 ± 8 1200 ± 300 230 ± 80 1.4 ± 0.3
–1.2 no above-background activity
a

Data corresponds to the average of at least three individual runs, and the errors correspond to the difference between the average and the replicate of greatest difference from the average. Activity is not reported if it did not exceed three times background in more than one replicate. The pH of the NaHCO3 solutions after purging with CO2 was 6.5 ± 0.1 and 8.7 ± 0.3 when purged with N2.

b

Potentials reported vs Ag/AgCl/KCl(1M).

Figure 3.

Figure 3

CPE experiments run in 0.5 M NaHCO3 and 1 M KCl; the concentration of CoMP11-Ac was 1 μM when present. CPE run in (a) CO2-saturated solution, pH = 6.5 ± 0.1 and (b) N2-saturated solution, pH = 8.7 ± 0.3. Samples were run under a pressure of 1 atm of the gas indicated (CO2 or N2). Potentials reported vs Ag/AgCl/KCl(1M).

There are a few reports of potential-dependent selectivity in molecular CO2RR catalysts.3840 For example, in a study of a group of Pd complexes, those complexes with more negative reduction potentials favor protonation of the metal to form a hydride (proposed to primarily lead to HER), whereas the complexes with less negative potentials favor protonation of Pd-coordinated CO2, yielding CO.39 In a more recent study of Pd molecular catalysts, the authors sought to improve the selectivity for CO2-to-CO conversion by increasing the overpotential of HER, which was achieved by installing proximal cations in the second sphere of the catalyst.40

In our case, we propose that the distinct behavior of CoMP11-Ac arises from a dependence of the CO2RR catalytic mechanism upon the applied potential. Because the ECo(II/I) of CoMP11-Ac is estimated to be −1.42 V,43 the catalytic reduction of CO2 to CO at −1.2 V must originate from a different catalysis-initiating redox event. We propose that CO2 binding to the catalyst is coupled to the electroreduction of the catalyst (Scheme 1). This phenomenon where the formation or cleavage of a bond between heavy (non-hydrogen) atoms is coupled to electron and/or proton transfer has been invoked in electrochemical systems before.23 One example comes from the analysis of the rate-limiting O–O bond cleavage in the electrochemical reduction of aliphatic peroxides. When an all-concerted (coupling of the bond cleavage to both electron and proton transfer) pathway is at play, the CV feature associated with the electrochemically driven O–O bond cleavage was found to be at significantly less negative potential than when a stepwise mechanism is favored.57 In another example, an intermolecular concerted proton–electron transfer bond cleavage was also found to be the rate-determining step in the catalytic reduction of CO2 to CO by electrogenerated Fe(0) porphyrins in an aprotic solvent.58 Finally, the catalytic electroreduction of alkyl cobalt porphyrins is an example where carbon–metal bond breaking/formation and proton transfer are proposed to be concerted.59

Scheme 1. CO2-to-CO Catalysis Mechanism Proposed to Operate at −1.2 V vs Ag/AgCl/KCl(1M).

Scheme 1

The diagonal blue arrow shows the CO2-addition-coupled electron transfer proposed to initiate catalysis; M/M corresponds to the formal Co(II/I) reduction.

Considering CO2 reduction by CoMP11-Ac, if a molecule of CO2 is appropriately prepositioned near the catalyst active site, it could bind the metal center in a manner concerted with electron transfer from the electrode to the Co(II) species ([M] in Scheme 1). Concerted pathways have the advantage of avoiding high-energy intermediates invoked in stepwise pathways.23,6064 However, this advantage can be counterbalanced by other kinetic penalties. This is particularly likely in reactions that involve the breaking or formation of bonds between heavy atoms.23 The low CV current at −1.2 V, relative to the feature at −1.4 V, may be due to this additional kinetic expenditure. Prepositioning the CO2 molecule prior to binding to the metal center may be enabled or enhanced by conformational changes occurring upon adsorption of the catalyst on the mercury electrode. Similar effects have been found to account for the enhanced CO2RR activity of Ni-cyclam using a mercury working electrode. Adsorption of Ni(cyclam) onto the mercury electrode is proposed to cause a flattening of the ligand, leading to enhanced CO desorption kinetics (often the rate-determining step in CO2 reduction to CO by molecular catalysts) and diminished catalyst deactivation via CO poisoning.45,46,6567

The more cathodic CV feature (at Eh = −1.4 V) develops at a potential near the Co(II/I) couple of CoMP11-Ac, suggesting that the dominant reaction mechanism at −1.4 V is initiated by the Co(II/I) reduction of the catalyst. Once the formal Co(I) species is formed, either CO2 addition or proton transfer from a proton donor HA to the catalyst may occur. Consequently, both CO2-to-CO and H2-evolution catalysis take place, resulting in lower selectivity for CO2 reduction at this more cathodic potential (Scheme 2).

Scheme 2. CO2-to-CO and HER Catalysis Mechanisms Proposed to Operate at −1.4 V vs Ag/AgCl/KCl(1M).

Scheme 2

M/M Corresponds to the Formal Co(II/I) Reduction that Initiates Catalysis.

Effects of CO2 Partial Pressure (PCO2)

In the mechanism outlined in Scheme 1 and proposed to be at play at −1.2 V, the catalysis-initiating redox event would entail a Nernstian equilibrium between Co(II)MP11-Ac (M in Scheme 1) and the CO2-bound one-electron reduced species, as depicted in eq 3.

graphic file with name cs2c03297_m003.jpg 3

Based on the Nernst equation for this process, we expect the half-wave potential (Eh) to shift anodically with increasing partial pressure of CO2 (PCO2) with a slope of 59.2 mV/decade, as shown in eqs 4 and 5. In these equations, n represents the number of electrons transferred (i.e., 1) and E°′ corresponds to the thermodynamic potential under standard conditions.

graphic file with name cs2c03297_m004.jpg 4
graphic file with name cs2c03297_m005.jpg 5

The CV feature near −1.2 V does not show a clear peak, hindering our ability to accurately determine Eh. Instead, we define Ei as the potential at which a constant current of 1.5 μA is reached. eq 5 can be then rewritten in terms of Ei to obtain eq 6. (Please note that with this approximation, the E°′ term loses any physical meaning.) This approach of using the potential at which a constant current is reached has been employed as a proxy for Eh when non-ideal voltammograms are encountered.68

graphic file with name cs2c03297_m006.jpg 6

We apply eq 6 to the voltammograms of CoMP11-Ac collected under different PCO2 (Figures 4a and S5) achieved using mixtures of CO2 and N2 with different known compositions. To avoid deviations between PCO2 and the concentration of CO2 in solution, we avoided the use of NaHCO3 as a buffer and instead used 3-(cyclohexylamino)-1-propanesulfonic acid (CAPS pKa 10.4); more information regarding the effects of buffers is provided in the next section. In Figure 4a, we can see that as PCO2 increases, the onset potential shifts anodically. A plot of Ei vs −log(PCO2) (Figure 4b) shows a slope of ∼66 mV, supporting the proposal that the binding of CO2 is coupled to the one-electron reduction of the catalyst, as outlined in Scheme 1.

Figure 4.

Figure 4

(a) Linear sweep voltammograms of 1 μM CoMP11-Ac in 50 mM CAPS, 0.1 M KCl, pH 6.0 ± 0.2 at 100 mV/s under different PCO2; the arrows indicate the direction of increasing PCO2. (b) Plot of Ei vs −log(PCO2) showing a Nernstian slope of ∼66 mV/decade.

Effects of Proton Donor

To further test our mechanistic proposals, we varied the pKa of the proton donor HA, which under our experimental conditions, we anticipate being the conjugate acid form of the buffer.43,44,69 It has been reported that the pKa of the proton donor has a large impact on CO2-reduction catalysis. Relatively strong Brønsted acids lead to fast metal hydride formation and subsequent protonation to yield H2, as shown in the lower portion of Scheme 2; to minimize this undesirable pathway, weak Brønsted acids are preferred regardless of whether the catalyst operates in an aprotic or protic solvent.3,5,17,70 A particularly relevant example is the case of a water-soluble iron-porphyrin catalyst that was shown to evolve only H2 when using formate (pKa 3.7) buffer, while an equimolar mixture of CO and H2 was obtained in phosphate-buffered solution (pKa 7.2, H2PO4).30 In the case of CoMP11-Ac, we have previously reported that the rate of HER in water decreases with increasing buffer pKa due to a slower proton transfer from the buffer acid donor to the formal Co(I) and that such proton transfer to the formal Co(I) species is rate-limiting for buffers of pKa > 7.7.43 We have also found that the sterics of the proton donor species impact the catalytic CV current arising from HER catalyzed by a cobalt porphyrin mini enzyme in water.69 When exploring the effects of buffer properties on the CO2RR of a Ni(cyclam) electrocatalyst in water, the authors concluded that charge density (i.e., charge and size) of the buffer acid species was the main factor impacting the catalytic activity.71 With these precedents in mind, we chose to study the CO2RR of CoMP11-Ac in the presence of three structurally related buffers as proton donors with different pKa values: 3-(cyclohexylamino)-1-propanesulfonic acid (CAPS, pKa 10.4), 3-(cyclohexylamino)-1-ethanesulfonic acid (CHES, pKa 9.3), and 3-morpholinopropane-1-sulfonic acid (MOPS, pKa 7.2; structures shown in Figure 5).

Figure 5.

Figure 5

CVs of 1 μM CoMP11-Ac in (a) 50 mM CAPS, pH 5.3 ± 0.1; (b) 50 mM CHES, pH 5.9 ± 0.1; and (c) 50 mM MOPS, pH 5.9 ± 0.1. For all CVs, [KCl] = 0.1 M and scan rate = 100 mV/s. Arrows in the CV traces indicate the scanning direction.

The CVs of CoMP11-Ac in CO2-saturated solutions containing either CAPS, CHES, or MOPS (Figure 5) exhibit two features, one that peaks around −1.5 V, present also in N2-saturated solution, and a more anodic wave that starts developing near −1.2 V, not present under N2. When the weakest Brønsted acid CAPS is present, the addition of CO2 leads to a significant enhancement in the catalytic CV current (Figure 5a); as the pKa of the buffer acid decreases, the enhancement seen under CO2 relative to N2 becomes less pronounced. For MOPS solutions, the catalytic peak current seen for CoMP11-Ac is similar under both CO2 and N2 (Figure 5c). The lower pKa of MOPS is proposed to facilitate proton transfer to the catalyst,43 yielding a higher CV current (faster catalysis) under both CO2 and N2. This result indicates possible enhancement of H2 and CO evolution, both of which are impacted by the availability of protons. The catalytic peak currents seen for CoMP11-Ac decrease as the pKa of the buffer present increases. We attribute this trend to the lower acidity of the buffer acid species (i.e., higher pKa) disfavoring the transfer of protons, thereby slowing catalysis.

To assess whether product distribution is sensitive to the proton donor pKa and the applied potential, we performed CPE experiments at both −1.4 and −1.2 V in solutions containing CAPS, CHES, or MOPS buffers; the results are summarized in Table 2 (see Tables S2–S4 for information on individual runs). For the CPE experiments at −1.4 V under N2 (Figures S6–S8), the total charge passed in the presence of MOPS buffer is significantly higher than in CHES and CAPS. The lower pKa of the conjugate acid form of MOPS allows for more evolved H2, leading to a higher charge buildup, consistent with the trends seen in CV above as well as prior work.43 At −1.4 V in all CAPS, CHES, and MOPS, H2 is the sole product detected under N2, with respective FE values of 83, 92, and 92%. The TON for H2 is over 40-fold higher for catalysis in the presence of MOPS compared to CAPS. Overall, the charge passed decreases with increasing buffer acid pKa, a finding that is consistent with previous studies of buffer effects on HER by CoMP11-Ac,43 as well as with similar observations made for other catalysts working in both aqueous and aprotic solvents.19,44,69,7275

Table 2. Summary of 2-h CPE Results for 1 μM CoMP11-Ac in 0.5 M of the Specified Buffera.

  buffer E (V)b FE(H2) (%) FE(CO) (%) TON(H2) TON(CO) QT (C)
CO2 CAPS (pKa 10.4) –1.4 29 ± 6 48 ± 10 280 ± 10 470 ± 10 0.9 ± 0.2
–1.2 5 ± 1 88 ± 11 80 ± 20 1500 ± 300 1.7 ± 0.6
CHES (pKa 9.3) –1.4 43 ± 9 57 ± 4 940 ± 30 1300 ± 300 2.2 ± 0.3
–1.2 6 ± 1 81 ± 2 250 ± 30 3500 ± 300 4.2 ± 0.4
MOPS (pKa 7.2) –1.4 63 ± 13 21 ± 5 4100 ± 500 1400 ± 500 6.4 ± 0.8
–1.2 8 ± 2 85 ± 2 1200 ± 100 12,000 ± 1000 14.1 ± 1.4
N2 CAPS (pKa 10.4) –1.4 83 ± 16 ∼0 500 ± 90 ∼0 0.6 ± 0.2
–1.2 no above-background activity
CHES (pKa 9.3) –1.4 92 ± 11 ∼0 2800 ± 100 ∼0 3.0 ± 1.1
–1.2 67 ± 5 ∼0 590 ± 100 ∼0 0.8 (0.1)
MOPS (pKa 7.2) –1.4 92 ± 6 ∼0 23,000 ± 2000 ∼0 24.9 ± 4.8
–1.2 98 ± 3 ∼0 5000 ± 900 ∼0 4.7 ± 0.6
a

Data shown corresponds to the average of at least three individual runs, and the errors correspond to the difference between the average and the replicate with the greatest difference from the average. Activity is not reported if it did not exceed three times background in more than one replicate. The pH of all MOPS, CHES, and CAPS solutions after purging with CO2 was 6.5 ± 0.2, and 7.2 ± 0.2 when purged with N2.

b

Potentials reported vs Ag/AgCl/KCl(1M).

For CPE experiments on CoMP11-Ac conducted at −1.4 V under CO2 (Figures S6–S8 and Tables S2–S4), both CO and H2 are produced with appreciable yields for all three buffers. In CAPS, the FEs for CO and H2 are 48 and 29%, respectively, while in MOPS, these quantities are 21 and 63%. Thus, the pKa of the buffer is found to impact the product distribution at −1.4 V, with the lowest-pKa buffer MOPS favoring H2 formation the most. This finding supports the proposed mechanism (Scheme 2) and is consistent with other observations on CO2RR in water.6,17,18,30,53,55,76,77 We propose that the stronger the acid, the more rapidly the Co(I) species is protonated, enhancing the generation of H2. Weaker acids (CAPS and CHES) protonate this species more slowly, allowing CO2 binding to the formal Co(I) active species and leading to more conversion of CO2 to CO. This model is consistent with previous work on CoMP11-Ac, in which more acidic buffers were found to promote fast HER catalysis and were proposed to protonate the formal Co(I) species more rapidly.43

When CPEs of CoMP11-Ac are carried out at −1.2 V under N2 in CAPS and CHES, activity is low, being comparable to the background in CAPS and barely above background for CHES (Figures S6–S8 and Tables S2–S4). In MOPS, at −1.2 V, H2 is the only product and is detected with 98% FE and TON of 4,900 after 2 hours. This result indicates that −1.2 V is too anodic relative to E(Co(II/I)) to support HER activity unless a relatively acidic proton source (here, MOPS) is present. Previous work on HER by CoMP11-Ac showed that the presence of an acidic proton donor (pKa < 7.7) gives rise to a kinetic shift in the CV, allowing catalysis to occur at −1.2 V.43 Results under CO2 reveal a sharp contrast. For CPEs of CoMP11-Ac at −1.2 V under CO2, the overall activity is significantly higher than under N2 (Figures S6–S8). FE(CO) is nearly the same for CoMP11-Ac in all three buffers: CAPS (88%), CHES (81%), and MOPS (85%), and FE(H2) also is nearly the same in CAPS (5%), CHES (6%), and MOPS (8%) (Tables 2 and S2–S4). The insensitivity of the product distribution at −1.2 V to buffer pKa supports our proposal that in the presence of CO2, catalysis is initiated by CO2 binding coupled to catalyst reduction, which avoids the accumulation of a formal Co(I) species, leading to almost exclusive CO formation regardless of the acidic strength of the proton donor (Scheme 1). In other words, the selectivity-determining step precedes any proton transfer from the buffer acids, favoring the formation of CO irrespective of the proton donor pKa. Also worth highlighting is the TON(CO) of 12,000 achieved at −1.2 V after a 2-h CPE in MOPS under CO2, which compares well with other molecular electrocatalysts operating in water.7882

An interesting trend seen in the CPEs of CoMP11-Ac in the presence of CO2 in all four buffers (CAPS, CHES, MOPS, and NaHCO3) is that the catalyst is not only more selective for CO2 reduction at the less cathodic potential of −1.2 V but also exhibits similar or higher TON(CO) (Tables 1 and 2). We have previously reported that CoMP11-Ac experiences partial deactivation during electrocatalytic HER.41 This is consistent with the lower FE seen in CPE at the more negative potential (−1.4 V) and with the shape of the CV in which the current rapidly drops after reaching its maximum value between −1.4 and −1.5 V, as well as the inverse peak feature seen in the return scan, which is consistent with reactivation.50 We propose that enhanced catalyst deactivation is responsible, at least in part, for the lower total charge passed and overall FE at −1.4 V, (particularly when compared to −1.2 V under CO2). The coupled mechanism outlined in Scheme 1 would allow for CO2 reduction catalysis to occur at potentials at which catalyst deactivation is minimal, yielding the higher charge passed for CoMP11-Ac at −1.2 V under CO2. Indeed, the CPE traces of CoMP11-Ac at −1.2 V after 2 hours remain linear, indicating that the catalyst is still active (Figures S6–S8). Furthermore, CoMP11-Ac under CO2 in CAPS displays minimal deviation from linearity in the charge vs time CPE trace in a 24-h CPE at −1.2 V, yielding a TON(CO) of 9300. The 24-h CPE of CoMP11-Ac under CO2 in MOPS reveals some loss of activity after ∼6 h, as the CPE trace levels off, yet this more acidic proton donor yields a TON(CO) of 32,000 in 24 hs (Table S5 and Figures S9–S13).

To determine whether the enhanced catalyst deactivation at −1.4 V is responsible for the lower selectivity for CO at this potential, we performed CPE experiments on CoMP11-Ac under CO2 at −1.4 V for 24 hours (Figure S11 and Table S5) and compared FE(H2) and FE(CO) to the results obtained after the 2-hour CPE under otherwise identical conditions (Tables 2 and S2). The overall FE is lower at 24 h (69%), as expected for a longer bulk electrolysis experiment (attributed to more catalyst degradation), but FE(CO) is similar at 24 h (58 ± 6%) and 2 h (48 ± 10%). Interestingly, FE(H2) is lower at 24 h (11 ± 6%) vs 2 h (29 ± 6%), which suggests that the CoMP11-Ac deactivation product is not a more active HER catalyst. Instead, this data suggests that the deactivation product may be generated within the HER mechanism of CoMP11-Ac.

Conclusions

CoMP11-Ac catalyzes the reduction of CO2 to CO in water with FE(CO) up to 88%, with better selectivity at −1.2 V compared to −1.4 V. The high faradic efficiency for CO production seen in CPE at −1.2 V is proposed to originate from a distinct mechanism initiated by CO2 addition coupled to the reduction of the catalyst, avoiding accumulation of a formal Co(I) species. The lower selectivity found at −1.4 V is proposed to arise from the Co(II/I) reduction initiating catalysis, as the formal Co(I) species can undergo either CO2 addition or protonation, where the latter enables HER. Altogether, at the lower applied overpotential, CoMP11-Ac shows higher selectivity toward CO2-to-CO conversion as well as enhanced catalyst longevity. These results demonstrate how applied potential and proton donor pKa act together to determine catalyst selectivity. An implication is that these factors may contribute to system selectivity in complex ways, requiring codesign when developing and optimizing catalytic systems.

Acknowledgments

This work was supported by the US Department of Energy, Office of Science, Office of Basic Energy Sciences, under Award DE-SC0002106. This material is based upon work supported by the National Science Foundation Graduate Research Fellowship Program under Grant No. (DGE-1939268). Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.2c03297.

  • Experimental details, additional electrochemical data, tables of CPE results (Tables S1–S5) (Figures S1–S19) (PDF)

Author Present Address

Department of Chemistry, Yale University, New Haven, Connecticut 06520-8107, United States

Author Contributions

J.L.A.-H. and A.A.S. contributed equally to this paper.

The authors declare no competing financial interest.

Supplementary Material

cs2c03297_si_001.pdf (1.2MB, pdf)

References

  1. National Academies of Sciences, Engineering, and Medicine . Gaseous Carbon Waste Streams Utilization: Status and Research Needs; The National Academies Press: Washington, DC, 2019; pp 63–96. [Google Scholar]
  2. Burkart M. D.; Hazari N.; Tway C. L.; Zeitler E. L. Opportunities and Challenges for Catalysis in Carbon Dioxide Utilization. ACS Catal. 2019, 9, 7937–7956. 10.1021/acscatal.9b02113. [DOI] [Google Scholar]
  3. 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]
  4. 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. F.; 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]
  5. Costentin C.; Robert M.; Savéant J.-M. Catalysis of the electrochemical reduction of carbon dioxide. Chem. Soc. Rev. 2013, 42, 2423–2436. 10.1039/C2CS35360A. [DOI] [PubMed] [Google Scholar]
  6. Shafaat H. S.; Yang J. Y. Uniting biological and chemical strategies for selective CO2 reduction. Nat. Catal. 2021, 4, 928–933. 10.1038/s41929-021-00683-1. [DOI] [Google Scholar]
  7. Schneider J.; Jia H.; Muckerman J. T.; Fujita E. Thermodynamics and kinetics of CO2, CO, and H+ binding to the metal centre of CO2 reduction catalysts. Chem. Soc. Rev. 2012, 41, 2036–2051. 10.1039/C1CS15278E. [DOI] [PubMed] [Google Scholar]
  8. Franco F.; Rettenmaier C.; Jeon H. S.; Roldan Cuenya B. Transition metal-based catalysts for the electrochemical CO2 reduction: from atoms and molecules to nanostructured materials. Chem. Soc. Rev. 2020, 49, 6884–6946. 10.1039/D0CS00835D. [DOI] [PubMed] [Google Scholar]
  9. Chen L.; Guo Z.; Wei X.-W.; Gallenkamp C.; Bonin J.; Anxolabéhére-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]
  10. Takeda H.; Cometto C.; Ishitani O. Marc Robert, M. Electrons, Photons, Protons and Earth-Abundant Metal Complexes for Molecular Catalysis of CO2 Reduction. ACS Catal. 2017, 7, 70–88. 10.1021/acscatal.6b02181. [DOI] [Google Scholar]
  11. 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]
  12. Fukuzumi S.; Lee Y.-M.; Ahn H. S.; Nam W. Mechanisms of catalytic reduction of CO2 with heme and nonheme metal complexes. Chem. Sci. 2018, 9, 6017–6034. 10.1039/C8SC02220H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Zhang B.; Sun L. Artificial photosynthesis: opportunities and challenges of molecular catalysts. Chem. Soc. Rev. 2019, 48, 2216–2264. 10.1039/C8CS00897C. [DOI] [PubMed] [Google Scholar]
  14. Jiang C.; Nichols A. W.; Machan C. W. A look at periodic trends in d-block molecular electrocatalysts for CO2 reduction. Dalton Trans. 2019, 48, 9454–9468. 10.1039/C9DT00491B. [DOI] [PubMed] [Google Scholar]
  15. Le J. M.; Bren K. L. Engineered Enzymes and Bioinspired Cataysts for Energy Conversion. ACS Energy Lett. 2019, 4, 2168–2180. 10.1021/acsenergylett.9b01308. [DOI] [Google Scholar]
  16. 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]
  17. Bhugun I.; Lexa D.; Savéant J.-M. Catalysis of the electrochemical reduction of carbon dioxide by iron(0) porphyrins: Synergystic effect of weak Brönsted acids. J. Am. Chem. Soc. 1996, 118, 1769–1776. 10.1021/ja9534462. [DOI] [Google Scholar]
  18. Bhugun I.; Lexa D.; Savéant J.-M. Ultraefficient selective homogeneous catalysis of the electrochemical reduction of carbon dioxide by an iron(0) porphyrin associated with a weak Brönsted acid co-catalyst. J. Am. Chem. Soc. 1994, 116, 5015–5016. 10.1021/ja00090a068. [DOI] [Google Scholar]
  19. Rountree E. S.; Martin D. J.; McCarthy B. D.; Dempsey J. L. Linear Free Energy Relationships in the Hydrogen Evolution Reaction: Kinetic Analysis of a Cobaloxime Catalyst. ACS Catal. 2016, 6, 3326–3335. 10.1021/acscatal.6b00667. [DOI] [Google Scholar]
  20. 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]
  21. Bhugun I.; Lexa D.; Savéant J.-M. Catalysis of the electrochemical reduction of carbon dioxide by iron(0) porphyrins. Synergistic effect of Lewis acid cations. J. Phys. Chem. A 1996, 100, 19981–19985. 10.1021/jp9618486. [DOI] [Google Scholar]
  22. Bernskoetter W. H.; Hazari N. Reversible Hydrogenation of Carbon Dioxide to Formic Acid and Methanol: Lewis Acid Enhancement of Base Metal Catalysts. Acc. Chem. Res. 2017, 50, 1049–1058. 10.1021/acs.accounts.7b00039. [DOI] [PubMed] [Google Scholar]
  23. Costentin C.; Robert M.; Savéant J.-M.; Tard C. Breaking Bonds with Electrons and Protons. Models and Examples. Acc. Chem. Res. 2014, 47, 271–280. 10.1021/ar4001444. [DOI] [PubMed] [Google Scholar]
  24. Häckl K.; Kunz W. Some aspects of green solvents. C. R. Chim. 2018, 21, 572–580. 10.1016/j.crci.2018.03.010. [DOI] [Google Scholar]
  25. Schneider C. R.; Shafaat H. S. An internal electron reservoir enhances catalytic CO2 reduction by a semisynthetic enzyme. Chem. Commun. 2016, 52, 9889–9892. 10.1039/C6CC03901D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. 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]
  27. Chen Z.; Chen C.; Weinberg D. R.; Kang P.; Concepcion J. J.; Harrison D. P.; Brookhart M. S.; Meyer T. J. Electrocatalytic reduction of CO2 to CO by polypyridyl ruthenium complexes. Chem. Commun. 2011, 47, 12607–12609. 10.1039/c1cc15071e. [DOI] [PubMed] [Google Scholar]
  28. Beley M.; Collin J. P.; Ruppert R.; Sauvage J. P. Nickel(II)-Cyclam: an Extremely Selective Electrocatalyst for Reduction of CO2 in Water. J. Chem. Soc., Chem. Commun. 1984, 1315–1316. 10.1039/c39840001315. [DOI] [Google Scholar]
  29. Taheri A.; Thompson E. J.; Fettinger J. C.; Berben L. A. An Iron Electrocatalyst for Selective Reduction of CO2 to Formate in Water: Including Thermochemical Insights. ACS Catal. 2015, 5, 7140–7151. 10.1021/acscatal.5b01708. [DOI] [Google Scholar]
  30. Costentin C.; Robert M.; Savéant J.-M.; Tatin A. Efficient and selective molecular catalyst for the CO2-to-CO electrochemical conversion in water. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 6882–6886. 10.1073/pnas.1507063112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Lide D. R.Handbook of Chemistry and Physics, 81st ed.; CRC Press, 2000; Vol. 122, pp 8–90. [Google Scholar]
  32. 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]
  33. Stratakes B. M.; Dempsey J. L.; Miller A. J. M. Determining the Overpotential of Electrochemical Fuel Synthesis Mediated by Molecular Catalysts: Recommended Practices, Standard Reduction Potentials, and Challenges. ChemElectroChem 2021, 8, 4161–4180. 10.1002/celc.202100576. [DOI] [Google Scholar]
  34. 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]
  35. Tomisaki M.; Kasahara S.; Natsui K.; Ikemiya N.; Einaga K. Switchable Product Selectivity in the Electrochemical Reduction of Carbon Dioxide Using Boron-Doped Diamond Electrodes. J. Am. Chem. Soc. 2019, 141, 7414–7420. 10.1021/jacs.9b01773. [DOI] [PubMed] [Google Scholar]
  36. Ju W.; Bagger A.; Hao G.-P.; Varela A. S.; Sinev I.; Bon V.; Roldan-Cuenya B.; Kaskel S.; Rossmeisl J.; Strasser P. Understanding activity and selectivity of metal-nitrogen-doped carbon catalysts for electrochemical reduction of CO2. Nat. Commun. 2017, 8, 944 10.1038/s41467-017-01035-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Zhu C.; Lan B.; Wei R.-L.; Wang C.-N.; Yang Y.-Y. Potential-Dependent Selectivity of Ethanol Complete Oxidation on Rh Electrode in Alkaline Media: A Synergistic Study of Electrochemical ATR-SEIRAS and IRAS. ACS Catal. 2019, 9, 4046–4053. 10.1021/acscatal.9b00138. [DOI] [Google Scholar]
  38. Weng Z.; Jiang J.; Wu Y.; Wu Z.; Guo X.; Materna K. L.; Liu W.; Batista V. S.; Brudvig G. W.; Wang H. Electrochemical CO2 Reduction to Hydrocarbons on a Heterogeneous Molecular Cu Catalyst in Aqueous Solution. J. Am. Chem. Soc. 2016, 138, 8076–8079. 10.1021/jacs.6b04746. [DOI] [PubMed] [Google Scholar]
  39. DuBois D. L. Development of Transition Metal Phosphine Complexes as Electrocatalysts for CO2 and CO Reduction. Comments Inorg. Chem. 1997, 19, 307–325. 10.1080/02603599708032743. [DOI] [Google Scholar]
  40. Barlow J. M.; Ziller J. W.; Yang J. Y. Inhibiting the Hydrogen Evolution Reaction (HER) with Proximal Cations: A Strategy for Promoting Selective Electrocatalytic Reduction. ACS Catal. 2021, 11, 8155–8164. 10.1021/acscatal.1c01527. [DOI] [Google Scholar]
  41. Kleingardner J. G.; Kandemir B.; Bren K. L. Hydrogen Evolution from Neutral Water under Aerobic Conditions Catalyzed by Cobalt Microperoxidase-11. J. Am. Chem. Soc. 2014, 136, 4–7. 10.1021/ja406818h. [DOI] [PubMed] [Google Scholar]
  42. Edwards E. H.; Jelusic J.; Chakraborty S.; Bren K. L. Photochemical hydrogen evolution from cobalt microperoxidase-11. J. Inorg. Biochem. 2021, 217, 111384 10.1016/j.jinorgbio.2021.111384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Alvarez-Hernandez J. L.; Sopchak A. E.; Bren K. L. Buffer pKa Impacts the Mechanism of Hydrogen Evolution Catalyzed by a Cobalt Porphyrin-Peptide. Inorg. Chem. 2020, 59, 8061–8069. 10.1021/acs.inorgchem.0c00362. [DOI] [PubMed] [Google Scholar]
  44. Alvarez-Hernandez J. L.; Han J. W.; Sopchak A. E.; Guo Y.; Bren K. L. Linear Free Energy Relationships in Hydrogen Evolution Catalysis by a Cobalt Tripeptide in Water. ACS Energy Lett. 2021, 6, 2256–2261. 10.1021/acsenergylett.1c00680. [DOI] [Google Scholar]
  45. Balazs G. B.; Anson F. C. The adsorption of Ni(cyclam)+ at mercury electrodes and its relation to the electrocatalytic reduction of CO2. J. Electroanal. Chem. 1992, 322, 325–345. 10.1016/0022-0728(92)80086-J. [DOI] [Google Scholar]
  46. Wu Y.; Rudshteyn B.; Zhainadarova A.; Froehlich J. D.; Ding W.; Kubiak C. P.; Batista V. S. Electrode-Ligand Interactions Dramatically Enhance CO2 Conversion to CO by the [Ni(cyclam)](PF6)2 Catalyst. ACS Catal. 2017, 7, 5282–5288. 10.1021/acscatal.7b01109. [DOI] [Google Scholar]
  47. Artero V.; Fontecave M. Solar fuels generation and molecular systems: is it homogeneous or heterogeneous catalysis?. Chem. Soc. Rev. 2013, 42, 2338–2356. 10.1039/C2CS35334B. [DOI] [PubMed] [Google Scholar]
  48. Lee K. J.; McCarthy B. D.; Dempsey J. L. On decomposition, degradation, and voltammetric deviation: the electrochemist’s field guide to identifying precatalyst transformation. Chem. Soc. Rev. 2019, 48, 2927–2945. 10.1039/C8CS00851E. [DOI] [PubMed] [Google Scholar]
  49. Widegren J. A.; Finke R. G. A review of the problem of distinguishing true homogeneous catalysis from soluble or other metal-particle hetereogeneous catalysis under reducing conditions. J. Mol. Catal. A: Chem. 2002, 198, 317–341. 10.1016/S1381-1169(02)00728-8. [DOI] [Google Scholar]
  50. Limoges B.; Savéant J.-M. Catalysis by immobilized redox enzymes. Diagnosis of inactivation and reactivation effects through odd cyclic voltammetric responses. J. Electroanal. Chem. 2004, 562, 43–52. 10.1016/j.jelechem.2003.07.035. [DOI] [Google Scholar]
  51. Nie W.; Tarnopol D. E.; McCrory C. C. L. Enhancing a Molecular Electrocatalyst’s Activity for CO2 Reduction by Simultaneously Modulating Three Substituent Effects. J. Am. Chem. Soc. 2021, 143, 3764–3778. 10.1021/jacs.0c09357. [DOI] [PubMed] [Google Scholar]
  52. Nie W.; Wang Y.; Zheng T.; Ibrahim A.; Xu Z.; McCrory C. C. L. Electrocatalytic CO2 Reduction by Cobalt Bis(pyridylmonoimine) Complexes: Effect of Ligand Flexibility on Catalytic Activity. ACS Catal. 2020, 10, 4942–4959. 10.1021/acscatal.9b05513. [DOI] [Google Scholar]
  53. Nie W.; McCrory C. C. L. Electrocatalytic CO2 reduction by a cobalt bis(pyridylmonoimine) complex: effect of acid concentration on catalyst activity and stability. Chem. Commun. 2018, 54, 1579–1582. 10.1039/C7CC08546J. [DOI] [PubMed] [Google Scholar]
  54. 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]
  55. Clark M. L.; Cheung P. L.; Lessio M.; Carter E. A.; Kubiak C. P. Kinetic and Mechanistic Effects of Bipyridine (bpy) Substituent, Labile Ligand, and Brønsted Acid on Electrocatalytic CO2 Reduction by Re(bpy) Complexes. ACS Catal. 2018, 8, 2021–2029. 10.1021/acscatal.7b03971. [DOI] [Google Scholar]
  56. Machan C. W.; Yin J.; Chabolla S. A.; Gilson M. K.; Kubiak C. P. Improving the Efficiency and Activity of Electrocatalysts for the Reduction of CO2 through Supramolecular Assembly with Amino Acid-Modified Ligands. J. Am. Chem. Soc. 2016, 138, 8184–8193. 10.1021/jacs.6b03774. [DOI] [PubMed] [Google Scholar]
  57. Costentin C.; Hajj V.; Robert M.; Savéant J.-M.; Tard C. Concerted heavy-atom bond cleavage and proton and electron transfers illustrated by proton-assisted reductive cleavage of an O–O bond. Proc. Natl. Acad. Sci. U.S.A. 2011, 21, 8559–8564. 10.1073/pnas.1104952108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Costentin C.; Drouet S.; Passard G.; Robert M.; Savéant J.-M. Proton-Coupled Electron Transfer Cleavage of Heavy-Atom Bonds in Electrocatalytic Processes. Cleavage of a C–O Bond in the Catalyzed Electrochemical Reduction of CO2. J. Am. Chem. Soc. 2013, 135, 9023–9031. 10.1021/ja4030148. [DOI] [PubMed] [Google Scholar]
  59. Costentin C.; Passard G.; Robert M.; Savéant J.-M. Concertedness in proton-coupled electron transfer cleavages of carbon–metal bonds illustrated by the reduction of an alkyl cobalt porphyrin. Chem. Sci. 2013, 4, 819–823. 10.1039/C2SC21788K. [DOI] [Google Scholar]
  60. Mayer J. M. Proton-Coupled Electron Transfer: A Reaction Chemist’s View. Annu. Rev. Phys. Chem. 2004, 55, 363–390. 10.1146/annurev.physchem.55.091602.094446. [DOI] [PubMed] [Google Scholar]
  61. Hammes-Schiffer S. Proton-Coupled Electron Transfer: Moving Together and Charging Forward. J. Am. Chem. Soc. 2015, 137, 8860–8871. 10.1021/jacs.5b04087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Weinberg D. R.; Gagliardi C. J.; Hull J. F.; Murphy C. F.; Kent C. A.; Westlake B. C.; Paul A.; Ess D. H.; McCafferty D. C.; Meyer T. J. Proton-Coupled Electron Transfer. Chem. Rev. 2012, 112, 4016–4093. 10.1021/cr200177j. [DOI] [PubMed] [Google Scholar]
  63. Costentin C.; Robert M.; Savéant J.-M. Update 1 of: Electrochemical Approach to the Mechanistic Study of Proton-Coupled Electron Transfer. Chem. Rev. 2010, 110, PR1–PR40. 10.1021/cr100038y. [DOI] [PubMed] [Google Scholar]
  64. Warren J. J.; Mayer J. M. Moving Protons and Electrons in Biomimetic Systems. Biochemistry 2015, 54, 1863–1878. 10.1021/acs.biochem.5b00025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Beley M.; Collin J. P.; Ruppert R.; Sauvage J. P. Electrocatalytic Reduction of CO2 by Ni(cyclam)2+ 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]
  66. Fujihira M.; Hirata Y.; Suga K. Electrocatalytic reduction of CO2 by nickel(II) cyclam. Study of the reduction mechanism on mercury by cyclic voltammetry, polarography and electrocapillarity. J. Electroanal. Chem. 1990, 292, 199–215. 10.1016/0022-0728(90)87336-I. [DOI] [Google Scholar]
  67. Bujno K.; Bilewicz R.; Siegfried L.; Kaden T. A. Effects of ligand structure on the adsorption of nickel tetraazamacrocyclic complexes and electrocatalytic CO2 reduction. J. Electroanal. Chem. 1998, 445, 47–53. 10.1016/S0022-0728(97)00603-7. [DOI] [Google Scholar]
  68. Wang D.; Groves J. T. Efficient water oxidation catalyzed by homogeneous cationic cobalt porphyrins with critical roles for the buffer base. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 11579–15584. 10.1073/pnas.1315383110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Le J. M.; Alachouzos G.; Chino M.; Frontier A. J.; Lombardi A.; Bren K. L. Tuning Mechanism through Buffer Acid Dependence of Hydrogen Evolution by a Cobalt Mini-enzyme. Biochemistry 2020, 59, 1289–1297. 10.1021/acs.biochem.0c00060. [DOI] [PubMed] [Google Scholar]
  70. Waldie K. M.; Ostericher A. L.; Reineke M. H.; Sasayama A. F.; Kubiak C. P. Hydricity of Transition-Metal Hydrides: Thermodynamic Considerations for CO2 Reduction. ACS Catal. 2018, 8, 1313–1324. 10.1021/acscatal.7b03396. [DOI] [Google Scholar]
  71. Schneider C. R.; Lewis L. C.; Shafaat H. S. The good, the neutral, and the positive: buffer identity impacts CO2 reduction activity by nickel(II) cyclam. Dalton Trans. 2019, 48, 15810–15821. 10.1039/C9DT03114F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Martin D. J.; Wise C. F.; Pegis M. L.; Mayer J. M. Developing Scaling Relationships for Molecular Electrocatalysis through Studies of Fe-Porphyrin-Catalyzed O2 Reduction. Acc. Chem. Res. 2020, 53, 1056–1065. 10.1021/acs.accounts.0c00044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Medina-Ramos J.; Oyesanya O.; Alvarez J. C. Buffer Effects in the Kinetics of Concerted Proton-Coupled Electron Transfer: The Electrochemical Oxidation of Glutathione Mediated by [IrCl6]2– at Variable Buffer pKa and Concentration. J. Phys. Chem. C 2013, 117, 902–912. 10.1021/jp3111265. [DOI] [Google Scholar]
  74. Jencks W. P. Requirements for General Acid-Base Catalysis of Complex Reactions. J. Am. Chem. Soc. 1972, 94, 4731–4732. 10.1021/ja00768a052. [DOI] [Google Scholar]
  75. Jencks W. P. General Acid-Base Catalysis of Complex Reactions in Water. Chem. Rev. 1972, 72, 705–718. 10.1021/cr60280a004. [DOI] [Google Scholar]
  76. Smieja J. M.; Sampson M. D.; Grice K. A.; Benson E. E.; Froehlich J. D.; Kubiak C. P. Manganese as a Substitute for Rhenium in CO2 Reduction Catalysts: The Importance of Acids. Inorg. Chem. 2013, 52, 2484–2491. 10.1021/ic302391u. [DOI] [PubMed] [Google Scholar]
  77. Ostericher A. L.; Waldie K. M.; Kubiak C. P. Utilization of Thermodynamic Scaling Relationships in Hydricity To Develop Nickel Hydrogen Evolution Reaction Electrocatalysts with Weak Acids and Low Overpotentials. ACS Catal. 2018, 8, 9596–9603. 10.1021/acscatal.8b02922. [DOI] [Google Scholar]
  78. Kramer W. W.; McCrory C. C. L. Polymer coordination promotes selective CO2 reduction by cobalt phthalocyanine. Chem. Sci. 2016, 7, 2506–2515. 10.1039/C5SC04015A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Smith P. T.; Benke B. P.; Cao Z.; Kim Y.; Nichols E. M.; Kim K.; Chang C. J. Iron Porphyrins Embedded into a Supramolecular Porous Organic Cage for Electrochemical CO2 Reduction in Water. Angew. Chem., Int. Ed. 2018, 57, 96884–99688. 10.1002/anie.201803873. [DOI] [PubMed] [Google Scholar]
  80. Morlanés N.; Takanabe K.; Rodionov V. Simultaneous Reduction of CO2 and Splitting of H2O by a Single Immobilized Cobalt Phthalocyanine Electrocatalyst. ACS Catal. 2016, 6, 3092–3095. 10.1021/acscatal.6b00543. [DOI] [Google Scholar]
  81. Zhang X.; Wu Z.; Zhang X.; Li L.; Li Y.; Xu H.; Li X.; Yu X.; Zhang Z.; Liang Y.; Wang H. Highly selective and active CO2 reduction electrocatalysts based on cobalt phthalocyanine/carbon nanotube hybrid structures. Nat. Comm. 2017, 8, 14675 10.1038/ncomms14675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Lin S.; Diercks C. S.; Zhang Y. B.; Kornienko N.; Nichols E. M.; Zho Y.; Paris A. R.; Kim D.; Yang P.; Yaghi O. M.; Chang C. J. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208–1213. 10.1126/science.aac8343. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

cs2c03297_si_001.pdf (1.2MB, pdf)

Articles from ACS Catalysis are provided here courtesy of American Chemical Society

RESOURCES