Next Article in Journal
The Potential of Simulating Energy Systems: The Multi Energy Systems Simulator Model
Next Article in Special Issue
A Perylenediimide-Based Zinc-Coordination Polymer for Photosensitized Singlet-Oxygen Generation
Previous Article in Journal
Energy Management System for Hybrid PV/Wind/Battery/Fuel Cell in Microgrid-Based Hydrogen and Economical Hybrid Battery/Super Capacitor Energy Storage
Previous Article in Special Issue
Molecular Dye-Sensitized Photocatalysis with Metal-Organic Framework and Metal Oxide Colloids for Fuel Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

FeI Intermediates in N2O2 Schiff Base Complexes: Effect of Electronic Character of the Ligand and of the Proton Donor on the Reactivity with Carbon Dioxide

Department of Chemical Sciences, University of Padova, Via Marzolo 1, 35131 Padova, Italy
*
Authors to whom correspondence should be addressed.
Energies 2021, 14(18), 5723; https://doi.org/10.3390/en14185723
Submission received: 29 July 2021 / Revised: 6 September 2021 / Accepted: 7 September 2021 / Published: 11 September 2021
(This article belongs to the Special Issue Advances in Molecular Artificial Photosynthesis)

Abstract

:
The characterization of competent intermediates of metal complexes, involved in catalytic transformations for the activation of small molecules, is an important target for mechanistic comprehension and catalyst design. Iron complexes deserve particular attention, due to the rich chemistry of iron that allows their application both in oxidation and reduction processes. In particular, iron complexes with tetradentate Schiff base ligands show the possibility to electrochemically generate FeI intermediates, capable of reacting with carbon dioxide. In this work, we investigate the electronic and spectroscopic features of FeI intermediates in five Fe(LN2O2) complexes, and evaluate the electrocatalytic reduction of CO2 in the presence of phenol (PhOH) or trifluoroethanol (TFE) as proton donors. The main findings include: (i) a correlation of the potentials of the FeII/I couples with the electronic character of the LN2O2 ligand and the energy of the metal-to-ligand charge transfer absorption of FeI species (determined by spectroelectrochemistry, SEC-UV/Vis); (ii) the reactivity of FeI species with CO2, as proven by cyclic voltammetry and SEC-UV/Vis; (iii) the identification of Fe(salen) as a competent homogeneous electrocatalyst for CO2 reduction to CO, in the presence of phenol or trifluoroethanol proton donors (an overpotential of 0.91 V, a catalytic rate constant estimated at 5 × 104 s−1, and a turnover number of 4); and (iv) the identification of sudden, ligand-assisted decomposition routes for complexes bearing a ketylacetoneimine pendant, likely associated with the protonation under cathodic conditions of the ligands.

Graphical Abstract

1. Introduction

Iron coordination compounds are intensively investigated for driving several chemical transformations, in particular involving oxidation/reduction processes. A key feature is the ample redox chemistry of the iron metal center that enables the formation of either high-valent FeIV or FeV oxo intermediates [1,2] or low-valent FeI or Fe0 derivatives [3,4,5,6,7] involved in the reductive transformation of protons, nitrogen, and carbon dioxide. Focusing in particular on CO2, the nature of the low-valent iron intermediate, together with the combination of suitable additives, have been shown to impact the selectivity of the process, with possible products being carbon monoxide, formate, or highly reduced methanol and methane [8].
Iron porphyrins, where iron is coordinated by a N4 tetradentate heme motif, are the most investigated class of coordination compounds for the reduction of CO2, and operate through the generation of a formal Fe0 state that binds CO2 and further promotes its conversion [9]. Conversely, iron complexes with tetradentate N2O2 Schiff base ligands have been less explored; examples have reported the use of 2-hydroxybenzene pendants on 1,10-phenanthroline (2,9-bis(2-hydroxyphenyl)-1,10-phenanthroline, dophen) [10], 2,2′-bipyridine scaffolds (6,6′-di(3,5-di-tert-butyl-2-hydroxybenzene)-2,2′-bipyridine, tBudhbpy) [11], and a recent report by some of us taking advantage of the N,N′-bis(salicylaldehyde)-1,2-phenylenediamine (salophen) ligand [12]. In Fe(salophen), the reactivity with CO2 involves an electrogenerated FeI intermediate, that in the presence of phenol as the proton donor, leads to the selective production of CO with appealing catalytic properties including an overpotential of 0.65 V and a turnover frequency up to 103 s−1 [12]; the nature and the concentration of the proton donor were found to play a key role in the catalytic process, in terms of product selectivity and of Fe(salophen) stability [12]. These findings, together with the ease of preparation of these complexes in gram scale, prompted us to investigate the reactivity of other Fe complexes with N2O2 ligands, focusing in particular on the characterization of the FeI intermediate and on its reactivity. Comprehension and rationalization of the reactivity of Fe complexes with N2O2 ligands are indeed important for catalyst design, aimed at the development of efficient CO2 conversion.
In this work, we consider four iron complexes with tetradentate N2O2 Schiff base ligands, bearing an additional chlorido ligand, FeIIICl(LN2O2), where LN2O2 is readily prepared by condensating ethylenediamine with salicylaldehyde, acetylacetone, or benzoylacetone. We will show that these complexes undergo two stepwise metal-centered reductions, leading to the formation of FeI intermediates. Reduction potentials of the FeII/I couples correlate with the DFT calculated electronic character of the LN2O2 and with spectroscopic features of the FeI intermediate. FeI species react with CO2, and the nature of the proton donor additive is crucial in promoting the transformation into CO with respect to competitive H2 evolution.

2. Materials and Methods

Synthetic procedures were accomplished by adapting literature procedures.
Cyclic voltammetry experiments were conducted with a three-electrode system controlled by a BASi EC Epsilon potentiostat-galvanostat. The working electrode was a glassy carbon disk electrode (BioLogic, nominal diameter 3 mm), the auxiliary electrode was a platinum electrode (BASi), and the reference electrode was an Ag/AgCl/NaCl (3 M) electrode; potentials were then referenced to the ferrocinium/ferrocene (Fc+/Fc) couple upon addition, at the end of each experiment session, of ferrocene to the analyte solutions as internal standard. 0.1 M tetraethylammonium tetrafluoroborate (Et4NBF4) was used as a supporting electrolyte.
Constant potential electrolysis experiments were performed with a Metrohm Autolab PGSTAT204 potentiostat-galvanostat controlled by Nova 2.1.4 software. The cell generally employed for preparative electrolysis was a custom-made, 6-necked, 2-compartment glass cell, with the two compartments being separated by a porous glass frit.
Gaseous product analysis was performed by gaschromatography equipped with a thermoconductimetric detector and a mass spectrometer. Gas samples of known volume were withdrawn from the headspace of the electrolysis cell by means of a gastight Hamilton syringe. Quantification of the gaseous species was achieved by external calibration of the instrument, upon the construction of a calibration curve by the injection of known volumes of pure gas.
1H-NMR analysis for formate detection followed the protocol reported by Machan et al., by treating 2 mL of the electrolysis solution with 2 mL of D2O, followed by washing twice with dichloromethane (formate is extracted in the D2O phase) [11].
UV/Visible absorption spectroelectrochemistry (SEC-UV/Vis) experiments were performed by employing a Varian Cary 50 Bio spectrophotometer, equipped with a 0.5 mm nominal optical path quartz spectroelectrochemistry cell (BASi EF-1362). The electrodes used were a platinum gauze working electrode, a platinum wire auxiliary electrode, and an Ag/AgCl/NaCl (3 M) reference electrode. The reference and auxiliary electrodes were immersed above the thin layer of solution electrolyzed by the working electrode.
DFT calculations were performed at the B3LYP/6–311+g(d,p) level of theory, with Gaussian16 software.
Further details are reported in Supplementary Information.

3. Results

3.1. Fe Complexes with Tetradentate N2O2 Ligands

The ligands employed in this work and the corresponding iron(III) coordination complexes, bearing an additional chlorido ligand (Figure 1), were prepared by adapting literature procedures [12,13,14]. The identity and purity of the synthesized compounds were confirmed by 1H and 13C-NMR (only for the ligands), and by electrospray ionization mass spectrometry (ESI-MS), elemental analysis, and UV/Vis spectrophotometry (see Supplementary Information).
The ligands employed in this work and the corresponding iron(III) coordination complexes, bearing an additional chlorido ligand (Figure 1), were prepared by adapting literature procedures [12,13,14]. The identity and purity of the synthesized compounds were confirmed by 1H and 13C-NMR (only for the ligands), and by electrospray ionization mass spectrometry (ESI-MS), elemental analysis, and UV/Vis spectrophotometry (see Supplementary Information).
In Fe(salen)Cl, iron is pentacoordinated and shows an asymmetric structure closer to a trigonal bipyramid with respect to a square pyramid, according to the τ geometric parameter [15,16]. Concerning N,N′-ehtylenebis(ketylacetonylideneimine) Iron(III) complexes, the N2O2 chelating group is almost planar, while the chlorido ligand occupies an axial position, in a distorted square planar coordination geometry [11,17]. In all cases, in the absence of additional high field ligands [17], Iron(III) is expected to have a d5 high spin (S = 5/2) configuration [11,17], and the complexes display ligand to metal charge transfer (LMCT) bands in the visible region (see Supplementary Information) [16]; absorptions in the UV region are due to intraligand charge-transfer transitions [18].

3.2. Electrochemical Properties and Spectroscopic Features of FeI Species

The electrochemical properties of Iron complexes were investigated by cyclic voltammetry (CV) in acetonitrile solutions, and are summarized in Table 1; Figure 2 (black traces under N2) reports the CV traces of Fe(salen)Cl and of Fe(beacen)Cl as representative cases, while other CV traces are reported in Supplementary Information (Figure S1). Although the redox-active nature of Schiff base ligands is well-established [19,20,21], common electrochemical features under a cathodic scan encompassing these coordination compounds are two one-electron waves, due to stepwise reduction processes of the iron center and attributed to FeIII/II and FeII/I couples, in which chlorido ligand binding equilibria are coupled to the electrochemical process driven by the working electrode potential (Figure 3) [11,12,22].
The electrochemical behavior of the complexes at the level of their FeIII/II couples, first described by Carré et al. [23], is typical of a quasi-reversible transformation. Specifically, all FeIII/II waves display ΔE in the range 80–100 mV, as summarized in Table 1 and reported in the Supplementary Information (Figure S2), coherently with competition between the coordinating solvent and the chloride anion for the apical coordination site in the FeIII and FeII states [11,12,22].
The second cathodic wave, encountered at more negative potentials, displays quasi-reversible features, depending on the Schiff base ligand coordinated to the iron ion. The structural and electronic analogies justifying the group categorization of the iron complexes herein described are manifest in the accessibility of a low valent formal FeI state.
The electronic properties of the Schiff base ligands and their direct influence on the redox potentials of the FeIII/II and FeII/I couples can be evaluated on the basis of the σ-donor strength of the N2O2 ligand [18,24]. Interestingly, in the series of the coordination complexes examined in this work, correlations are observed between the FeIII/II and FeII/I reduction potentials and the energy of the highest occupied molecular orbital of the ligands (Figure S3) involved in σ-donation (HOMO-2), as predicted by DFT calculations at the B3LYP/6–311+g(d,p) level of theory: Slopes of −0.40 ± 0.04 and of −0.84 ± 0.11 V/eV are observed for the FeIII/II and FeII/I couples, respectively (see Figures S4 and S5 in Supplementary Information). Notably, the almost double slope value observed for the FeII/I couples with respect to the FeIII/II ones indicates a higher impact of the nature of the ligand on the reduction potentials of the former couple. This could be ascribed to the fact that the FeIII→FeII transition is expected to be charge neutral upon loss of the Cl ligand, while the FeII→FeI transition is expected to generate a negative charge at the complex (Figure 3). The symmetry of the metal orbitals involved in the redox transformation is also important to support the observed changes in the redox properties caused by ligand electronic effects [18,25].
The formation of an FeI intermediate was previously corroborated by electron paramagnetic resonance (EPR) spectroscopy for Fe(salophen) [12]. Further confirmation of the electrochemical generation of the FeI species comes from spectroelectrochemistry in the UV/Vis region (SEC-UV/Vis). Figure 4 reports the absorption traces of Fe(salen)Cl and Fe(beacen)Cl as representative examples, while the other iron complexes are shown in Figure S6 in Supplementary Information. Upon application of the potentials associated with the second reduction waves (−2.21 and −2.24 V vs. Fc+/Fc for Fe(salen) and Fe(beacen), respectively), consistent UV/Vis spectra changes are observed (Figure 4 top), and in particular: (i) Bleaching of the initial absorption of the FeIII species (464 and 311 nm for Fe(salen)Cl and 492 and 293 nm for Fe(beacen)Cl), attributed to ligand-to-metal charge transfer bands, LMCT; and (ii) rising of new bands (344 nm for Fe(salen) and 354 nm for Fe(beacen)), attributed to metal-to-ligand charge transfer bands (MLCT) from the electron-rich FeI state.
Two further points are worthy of mention:
(i)
In the case of the Fe(acacsalen)Cl, two absorption features rise upon generation of FeI at 308 and 379 nm, respectively (see Figure S6); these two MLCT components are expected on the basis of the peculiar character of the asymmetric acacsalen ligand, bearing both the acetylacetoneimine and salicylideneimine pendants. Incidentally, the two absorption maxima observed for FeI(acacsalen) are similar to those observed separately for FeI(acacen) (288 nm) and for FeI(salen) (354 nm).
(ii)
The energy of the MLCT band observed for the FeI intermediates correlates linearly with the redox potential of the FeII/I couple; the trend shows that the more negative the potential of the FeII/I couple, the higher the energy of the MLCT band (absorption shifted towards the blue region of the spectrum, Figure S7).

3.3. Reactivity of FeI towards Carbon Dioxide

In the presence of carbon dioxide, the voltammetric traces of the iron complexes display intense, irreversible peak-shaped waves at the level of the FeII/I couple (see the iCO2/iN2 parameter summarized in Table 1, where iCO2 is the peak current of the wave under CO2, and iN2 is the peak current of the one-electron cathodic wave of the FeII/I couple under dinitrogen atmosphere). The features of the wave observed under a CO2 atmosphere are indicative of an electrocatalytic event involving the active FeI intermediate generated at the electrode, as evident in Figure 2 for Fe(salen)Cl and Fe(beacen)Cl as representative cases (see Figure S1 for the other complexes). In the case of Fe(beacen)Cl, the wave observed in the presence of CO2, is composed of two contributions. A pre-wave, observed at −2.20 V vs. Fc+/Fc (positively shifted by 80 mV with respect to the cathodic FeII/I peak under N2 atmosphere), is followed by the more intense wave peaking at Ep = −2.58 V vs. Fc+/Fc. Such an observation is likely ascribable to two different reduction pathways—associated with different reduced intermediates—of which the second, generated at more cathodic potentials, is involved in a faster catalytic reaction [26]. The first cathodic wave observed in the presence of CO2 could therefore be reasonably attributed to a slower process, proceeding through an adduct obtained as the irreversible association between CO2 and the reduced FeI complex, as previously observed in the case of an iron quaterpyridine catalyst [5]. Attempts to estimate association parameters between FeI and CO2 are hampered by the ill-defined nature of the pre-wave and the underlying current associated with the catalytic process, occurring at potentials close to the first process. We anticipate that in the presence of proton donors (vide infra), the subsequent catalytic current will be enhanced at this potential, most likely due to a protonation-first pathway of the reduced adduct, promoting the catalytic transformation [27].
Reactivity of the FeI species with CO2 was further supported by two complementary experiments:
(i)
SEC-UV/Vis analysis in the presence of CO2, where the diagnostic MLCT features of the FeI intermediate are blue-shifted with respect to those observed under dinitrogen (Figure 4 bottom). In particular, the absorption maximum shifts from 354 to 345 nm for Fe(salen) and from 344 nm to 332 nm for Fe(beacen); in the case of Fe(salophen), the MLCT shifted from 405 nm to 385 nm in the presence of CO2 [12]. A spectroelectrochemical analysis in the infrared region (SEC-IR) failed to reveal absorption features ascribable to stretching features of iron-carbonyl intermediates, as was observed in the case of Fe(salophen), ref. [12]. This could be ascribed to an enhanced reactivity of such intermediates, although the interference of platinum working electrode at the negative operative in our setup should be also considered.
(ii)
Isolation of FeI species upon constant potential electrolysis under inert atmosphere, followed by the addition of carbon dioxide (in the absence of applied potential), that led to an immediate color change of the solution (see pictures in Figure S8 in Supplementary Information). In the case of Fe(salophen), the reactivity of FeI with CO2 under analogous conditions was associated with a redox process involving the oxidation of FeI to FeIII by CO2, as supported by EPR evidence [12]. Notably, no reduced products of CO2 were detected under these conditions (in particular, carbon monoxide), suggesting that a further reduction of the FeI−CO2 adduct is required in order to close the cycle and release the products (vide infra).
The electrocatalytic process impacts the behavior observed in the backward scan. Indeed, the anodic trace recorded in the presence of CO2 displays an abatement of the FeII/III oxidation wave. In the case of Fe(salen)Cl, a new anodic peak (Ep = −0.93 V vs. Fc+/Fc) arises, 210 mV more negative than the FeII/III oxidation peak (Figure 2). These observations are ascribable to chemically irreversible transformations of the complexes induced by the applied potential in the presence of CO2 acting as a substrate. These processes might involve the loss or exchange of apical ligands in the coordination sphere of the iron ion, most likely involving CO [3,12,28,29]. Interestingly, a recovery of the reversibility of the FeIII/II couple in CV scans is observed in the presence of proton donors (see Section 3.4), which can facilitate the conversion of reduced intermediates and the regeneration of the parent form of the complex.

3.4. Effect of Proton Donors and Electrolysis

In order to evaluate and characterize the putative electrocatalytic reduction of CO2 in the presence of the iron complexes, the use of a proton donor adjuvant should be considered [3,27,28,29,30]. Following the indication of a previous screening conducted with Fe(salophen), phenol (PhOH, pKa = 29 in acetonitrile [29]) was considered, since it provided high selectivity for CO formation, also supported by 13CO2 labelling experiments [12]. Trifluoroethanol was also evaluated (TFE, pKa = 35.4 in acetonitrile [27,30], although a value of 25.1 was estimated in the presence of CO2 due to the reaction of the trifluoroethoxide conjugate base with CO2 to give the CF3CH2OCO2 carbonate [30]) since it was recently used in combination with Mn-based CO2 reduction catalysts [27]. The concentration of 0.3 M for TFE was selected as the one providing the maximum current in a CV screening with Fe(acacen)Cl.
The effect of the proton donors in the presence of CO2 was first investigated by CV (see representative traces in Figure 5 for Fe(beacen)Cl and Fe(salen)Cl; the CV traces for the other complexes are reported in Figures S9–S11), by choosing 0.3 M and 0.5 M as the concentrations of TFE and PhOH, respectively [12]. The major effect is a further current increase of the cathodic wave at the level of the FeII/I couple, as indicated by the iCO2/iN2 ratio (Table 2). This is coherent with a more efficient catalytic process, highlighting the beneficial role of mild proton donors in facilitating the further evolution of the putative Fe–CO2 intermediate upon its stabilization by hydrogen bonding and stepwise protonation [27,28]. Notably, the presence of trifluoroethanol led to the observation of the FeII→FeIII re-oxidation wave in the backward scans, likely indicative of a more efficient cycle, with a lower impact on the chemical stability of the catalytically active molecular species (vide supra).
Constant potential electrolysis (CPE) was then performed in order to identify and quantify the products associated with the cathodic waves, and to evaluate the stability of the systems. Three general considerations can be summarized as follows:
(i)
Carbon monoxide (CO) and hydrogen (H2) were the sole product identified; NMR analysis according to the procedure reported by Nichols et al. [11] did not reveal the production of formate (See Supplementary Information and Figure S14);
(ii)
In all cases, the overall Faradaic yield of the process for CO and H2 production is significantly lower than the ideal value; this is ascribed to both the need for pre-reducing the FeIII to the active FeI state, and the consumption of reducing equivalents by the iron complexes leading to their inactivation and decomposition, as demonstrated by the drop of electrolysis current over time (Figure S15) and by the marked changes in the UV/Vis spectra after electrolysis (Figure S16). Non-quantitative Faradaic yields are not unusual in the electrochemical reduction of CO2 with coordination complexes [5,12,31];
(iii)
The use of phenol as the proton donor leads to a marked impact on the selectivity of the process depending on the iron complex. In particular, while high selectivity for CO is observed in the case of Fe(salophen) and Fe(Salen), H2-oriented selectivity is observed for Fe(acacen), Fe(acacsalen), and Fe(beacen), as shown in Table 2 and Figure 6, i.e., for the species where the FeI intermediate is generated at more negative potentials and is thus expected to be more prone to a direct reaction with proton donors.

4. Discussion

From a catalytic perspective, Fe(salen)Cl appears the most promising candidate in the reduction of CO2. Fe(salen)Cl allowed the maintenance of a selectivity >98% along with electrolysis after 25 and 20 electrons passed per Iron center with PhOH and TFE, respectively, reaching a turnover number for CO production of 4.6 and 2.6 with PhOH and TFE, respectively. Although still limited, the TON values confirm the possibility for Fe(salen) to operate catalytically. For the sake of comparison, a TON of 8 and a Faradaic yield of 48% for CO were found under electrochemical conditions for one of the most recently investigated catalyst for reduction of CO2 to CO, i.e., Fe(qpy) (qpy = 2,2′:6′,2″:6″,2-quaterpyridine) [5,6]. Three TONs were registered for Fe(salophen) with PhOH [12]. The slightly higher stability of Fe(salen) with respect to Fe(salophen) could be associated with the more negative potential required for the reduction of the imine bond [32], which can be the origin of demetalation, responsible for the nucleation of Fe clusters [33] and the electrodeposition of Fe(0) nanoparticles [12,26,34]. CO2 reduction occurring through a molecular pathway is supported by the lack of any activity of the working electrode at the end of the CPE experiments. Indeed, the working electrode used in “unpolished tests” (i.e., CPE experiments in which the initial surface state of the working electrode was not restored run in electrolyte solutions without the catalyst, ref [26], pass only a limited amount of charge, displaying negligible activity towards proton reduction and no activity towards CO2 reduction. Eventual deposition of electroactive iron particles on the carbon electrode would instead produce an electrode prone to enhanced H2 evolution in the presence of proton donors, refs [12,34]. It is also worth to mention that in this latter case CO2 reduction to CH4 could reasonably be observed, with FE values in the 1–3% range, see refs [12,34]. A subsequent electrolysis on a Fe(salen) solution previously employed (25 electrons passed per iron centre, with phenol proton donor) displayed a lower current (<10% with respect to the first electrolysis) and negligible activity toward CO2 reduction. These results are coherent with Fe(salen) Faradaic decomposition side-processes to inert species. The robustness of the imine bond and the stabilization of the reduced FeI intermediate are thus the main target to improve operative stability. It is also worth mentioning that catalyst stability is extremely dependent on operative conditions (i.e., electrochemical or photochemical, see refs. [5,6]). In some cases, heterogenization of molecular catalysts was also shown to lead to improvements in stability.
Benchmarking of Fe(salen)Cl should also consider overpotential (η) and the rate constant (kcat), whose estimation can be directly determined by the CV analysis [35]. η was determined from the difference between the peak potential of the catalytic wave and the standard reduction potential associated with the CO2/CO evaluated in the same conditions (E0 (CO2/CO) = −1.34 vs. Fc+/Fc) [12,29]: In the presence of 0.5 M phenol, η results in 0.91 V, which is significantly higher than the one observed for Fe(salophen) under similar conditions, η = 0.65 V, due to the increased electron-donating character of the salen ligand that requires a more negative potential to generate the FeI species. Concerning kcat, given the impossibility of reaching an “S-shaped” wave typical of pure kinetic conditions even operating at high scan rates (up to 16 Vs−1), foot-of-the-wave analysis (FOWA) was conducted on the cyclic voltammetry (see treatment in Supplementary Information, and Figure S17), providing an estimation of kcat 5 × 104 s−1 (under the same conditions, Fe(salophen) exhibited 1 × 103 s−1) [12]. These key performance indicators are summarized in the Catalytic Tafel plot (Figure S18).
A final point of discussion deals with the initial CO vs. H2 selectivity observed in CPE experiments. The selectivity for CO of Fe(salen) was supported by the persistence of the spectroscopic features of the FeI intermediate along with SEC-UV/Vis experiments in the presence of both proton donors employed (Figure 7). This supports the limited reactivity of FeI(salen) with PhOH and TFE, thus favoring its reactivity with CO2 and supporting the observed CO2-to-CO reaction pathway (Table 2 and Figure 6). Further evidence is provided by the CV traces of Fe(salen)Cl in the presence of PhOH and TFE, where no current discharges attributable to H2 production at the level of the FeII/I couple are observed (Figures S10 and S11).
Conversely, for Fe(acacen)Cl, Fe(beacen(Cl), and Fe(acacsalen)Cl, the nature of the proton donor has an impact on the observed products [27], and in particular on the initial CO vs. H2 selectivity (Figure 6). While the use of TFE favors CO formation (selectivity 96, >99, and 73% for Fe(acacen)Cl, Fe(beacen(Cl), and Fe(acacsalen)Cl, respectively), the presence of phenol switches the process towards H2 production in the early stage of the experiment (associated with a drop in CO vs. H2 selectivity: 28, 41, <1% for Fe(acacen)Cl, Fe(beacen(Cl), and Fe(acacsalen)Cl, respectively, see Table 2). These results suggest direct reactivity of electrogenerated FeI species in this series with phenol, as confirmed by CV (rise in the cathodic current at the level of the FeII/I couple in the presence of phenol under N2 atmosphere, Figure S10) and by SEC-UV/Vis experiments. Indeed, the electrogeneration of the FeI intermediate under N2 atmosphere in the presence of phenol for Fe(beacen)Cl is a representative case (Figure S19), the diagnostic feature of FeI species is significantly abated, while new absorptions rise in the 400–800 nm region. Conversely, in the presence of TFE, a similar spectroscopic outcome with respect to the one previously discussed in the absence of proton donors is observed (rising of the MLCT band at 340 nm, see Figure 4 and related discussion).
The reactivity of FeI(beacen) with phenol can potentially involve both the iron center and the ligand. The H2 evolution observed for the β-diketyl-derived complexes in the presence of phenol could possibly be associated with the formation of iron hydride intermediates [11], and their further reaction with a second phenol equivalent; metal hydrides with sufficient hydricity [36] could also be responsible for formate production [37] (not detected in this study). Iron hydride generation might also be assisted by the previous protonation of the methine site, ultimately producing a hydride-transfer relay in the ligand scaffold [38]. Indeed, protonation of the imine group or of the methine carbon in β position to the imine and enolate can be considered under cathodic conditions [11,19], and both processes could account for the Faradaic decomposition of the catalysts; the presence of the methine group seems likely responsible for the faster complex degradation under electrolysis conditions for Fe(beacen), Fe(acacen), and Fe(acacsalen), regardless of the proton donor nature (the current drops after three electrons passed per iron center (Figure S15) and marked UV/Vis change (Figure S16)).

5. Conclusions

We have reported an electrochemical investigation of four FeIII(LN2O2)Cl complexes (LN2O2 is a tetradentate N2O2 Schiff base ligand), and their potential application in the reduction of carbon dioxide. The main results can be summarized as follows:
(i)
The FeIII(LN2O2)Cl complexes display two metal-based reductions, involving FeIII/II and FeII/I couples; the potential associated with the FeII/I couple, relevant to reactivity with CO2, correlates with the electronic character of the LN2O2 (in terms of the energy of the highest occupied σ-donating orbital) and with the energy of the metal-to-ligand charge transfer absorption of the FeI intermediate, determined by SEC-UV/Vis.
(ii)
The FeI intermediates react with CO2, as proven by CV and SEC-UV/Vis investigation.
(iii)
For Fe(salen)Cl, in the presence of phenol or trifluoroethanol proton donors, the process is associated with the selective reduction of CO2 to CO (no H2 and formate are detected along with the electrolysis); in the presence of 0.5 M phenol, key performance indicators are an overpotential of 0.91 V, a catalytic rate constant of 5 × 104 s−1, and a turnover number of 4. CO2 reduction occurs through a homogeneous route, and the transformation of Fe(Salen) into an electrochemically inert species occurs.
(iv)
In the case of Fe(acacen)Cl, Fe(beacen(Cl), and Fe(acacsalen)Cl, the production of CO is observed only with TFE proton donor, while phenol leads to the evolution of H2 from the early stage of electrolysis. In all cases, the electrolysis current drops suddenly after three electrons passed per iron center, indicating a higher instability of these species, likely associated with the protonation under cathodic conditions of the ketylacetoneimine pendant of the ligands.
The structure/reactivity correlations involving active low-valent Fe intermediates and the indications on the selectivity and stability of the coordination complexes under cathodic conditions and depending on the proton donor nature can be valuable in the design of more efficient and more robust catalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/en14185723/s1, including detailed experimental procedures and supplementary Figures S1–S19 cited in the main text.

Author Contributions

Conceptualization, R.B. and A.S.; methodology, R.B., D.C., F.C. and A.S.; formal analysis, R.B., D.C., F.C. and A.S.; investigation, R.B., D.C. and F.C.; writing, R.B. and A.S.; supervision, A.S.; project administration, A.S.; funding acquisition, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fondazione Cassa di Risparmio Padova e Rovigo (CaRiPaRo), grant Synergy within the call Ricerca Scientifica di Eccellenza 2018.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Acknowledgments

We thank Mauro Meneghetti and Stefano Mercanzin for the invaluable support in the construction of the gas-tight electrolysis cell and for technical support, and Loris Calore for the prompt support for elemental analysis of the compounds.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Oliveira, F.T.; Chanda, A.; Banerjee, D.; Shan, X.; Mondal, S.; Que, L.; Bominaar, E.L.; Münck, E.; Collins, T.J. Chemical and spectroscopic evidence for an FeV-oxo complex. Science 2007, 315, 835–838. [Google Scholar] [CrossRef] [PubMed]
  2. Bell, S.R.; Groves, J.T. A highly reactive P450 model compound I. J. Am. Chem. Soc. 2009, 131, 9640–9641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. 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. [Google Scholar] [CrossRef]
  4. Chalkley, M.J.; Drover, M.W.; Peters, J.C. Catalytic N2-to-NH3 (or -N2H4) Conversion by Well-Defined Molecular Coordination Complexes. Chem. Rev. 2020, 120, 5582–5636. [Google Scholar] [CrossRef]
  5. Cometto, C.; Chen, L.; Lo, P.K.; Guo, Z.; Lau, K.C.; Anxolabéhère-Mallart, E.; Fave, C.; Lau, T.C.; Robert, M. Highly Selective Molecular Catalysts for the CO2-to-CO Electrochemical Conversion at Very Low Overpotential. Contrasting Fe vs. Co Quaterpyridine Complexes upon Mechanistic Studies. ACS Catal. 2018, 8, 3411–3417. [Google Scholar] [CrossRef]
  6. Guo, Z.; Cheng, S.; Cometto, C.; Anxolabéhère-Mallart, E.; Ng, S.-M.; Ko, C.-C.; Lu, G.; Chen, L.; Robert, M.; Lau, T.-C. Highly Efficient and Selective Photocatalytic CO2 Reduction by Iron and Cobalt Quaterpyridine Complexes. J. Am. Chem. Soc. 2016, 138, 9413–9416. [Google Scholar] [CrossRef] [PubMed]
  7. Bhunia, S.; Rana, A.; Hematian, S.; Karlin, K.D.; Dey, A. Proton Relay in Iron Porphyrins for Hydrogen Evolution Reaction. Inorg. Chem. 2021. [Google Scholar] [CrossRef]
  8. Bonetto, R.; Crisanti, F.; Sartorel, A. Carbon Dioxide Reduction Mediated by Iron Catalysts: Mechanism and Intermediates That Guide Selectivity. ACS Omega 2020, 5, 21309–21319. [Google Scholar] [CrossRef] [PubMed]
  9. Costentin, C.; Robert, M.; Savéant, J.M. Current Issues in Molecular Catalysis Illustrated by Iron Porphyrins as Catalysts of the CO2-to-CO Electrochemical Conversion. Acc. Chem. Res. 2015, 48, 2996–3006. [Google Scholar] [CrossRef]
  10. Pun, S.N.; Chung, W.H.; Lam, K.M.; Guo, P.; Chan, P.H.; Wong, K.Y.; Che, C.M.; Chen, T.Y.; Peng, S.M. Iron(I) complexes of 2,9-bis(2-hydroxyphenyl)-1,10-phenanthroline (H2dophen) as electrocatalysts for carbon dioxide reduction. X-ray crystal structures of [Fe(dophen)Cl]2·2HCON(CH3)2 and [Fe(dophen)(N-MeIm)2]ClO4 (N-MeIm = 1-methylimidazole). J. Chem. Soc. Dalt. Trans. 2002, 575–583. [Google Scholar] [CrossRef]
  11. Nichols, A.W.; Chatterjee, S.; Sabat, M.; MacHan, C.W. Electrocatalytic Reduction of CO2 to Formate by an Iron Schiff Base Complex. Inorg. Chem. 2018, 57, 2111–2121. [Google Scholar] [CrossRef]
  12. Bonetto, R.; Altieri, R.; Tagliapietra, M.; Barbon, A.; Bonchio, M.; Robert, M.; Sartorel, A. Electrochemical Conversion of CO2 to CO by a Competent FeI Intermediate Bearing a Schiff Base Ligand. ChemSusChem 2020, 13, 4111–4120. [Google Scholar] [CrossRef] [PubMed]
  13. Nishida, Y.; Oshio, S.; Kida, S. Synthesis and Magnetic properties of Iron(III) complexes with several quadridentate Schiff bases. Bull. Chem. Soc. Jpn. 1977, 50, 119–122. [Google Scholar] [CrossRef]
  14. Cisterna, J.; Artigas, V.; Fuentealba, M.; Hamon, P.; Manzur, C.; Hamon, J.R.; Carrillo, D. Pentacoordinated chloro-iron(III) complexes with unsymmetrically substituted N2O2 quadridentate schiff-base ligands: Syntheses, structures, magnetic and redox properties. Inorganics 2018, 6, 5. [Google Scholar] [CrossRef] [Green Version]
  15. Liang, Y.; Duan, R.L.; Hu, C.Y.; Li, L.L.; Pang, X.; Zhang, W.X.; Chen, X.S. Salen-iron complexes: Synthesis, characterization and their reactivity with lactide. Chinese J. Polym. Sci. 2018, 36, 185–189. [Google Scholar] [CrossRef]
  16. Cozzolino, M.; Leo, V.; Tedesco, C.; Mazzeo, M.; Lamberti, M. Salen, salan and salalen iron(III) complexes as catalysts for CO2/epoxide reactions and ROP of cyclic esters. Dalt. Trans. 2018, 47, 13229–13238. [Google Scholar] [CrossRef]
  17. Wang, X.; Pennington, W.T.; Ankers, D.L.; Fanning, J.C. Comparative crystal structure examination of some iron(III) quadridentate schiff base complexes. Polyhedron 1992, 11, 2253–2264. [Google Scholar] [CrossRef]
  18. Böttcher, A.; Takeuchi, T.; Hardcastle, K.I.; Meade, T.J.; Gray, H.B.; Cwikel, D.; Kapon, M.; Dori, Z. Spectroscopy and Electrochemistry of Cobalt(III) Schiff Base Complexes. Inorg. Chem. 1997, 36, 2498–2504. [Google Scholar] [CrossRef] [Green Version]
  19. Isse, A.A.; Gennaro, A.; Vianello, E. Electrochemical reduction of Schiff base ligands H2salen and H2salophen. Electrochim. Acta 1997, 42, 2065–2071. [Google Scholar] [CrossRef]
  20. Chirik, P.J.; Wieghardt, K. Radical ligands confer nobility on base-metal catalysts. Science 2010, 327, 794–795. [Google Scholar] [CrossRef] [PubMed]
  21. Udugala-Ganehenege, M.Y.; Dissanayake, N.M.; Liu, Y.; Bond, A.M.; Zhang, J. Electrochemistry of nickel(II) and copper(II) N,N′-ethylenebis(acetylacetoniminato) complexes and their electrocatalytic activity for reduction of carbon dioxide and carboxylic acid protons. Transit. Met. Chem. 2014, 39, 819–830. [Google Scholar] [CrossRef]
  22. Ueda, T.; Inazuma, N.; Komatsu, D.; Yasuzawa, H.; Onda, A.; Guo, S.X.; Bond, A.M. Comparison of chemical interactions with Li+ and catalytic reactivity of electrochemically generated [FeICl(L)]2- and [CoI(L)]- complexes (L = salen or salophen). Dalt. Trans. 2013, 42, 11146–11154. [Google Scholar] [CrossRef] [PubMed]
  23. Costes, J.P.; Tommasino, J.B.; Carré, B.; Soulet, F.; Fabre, P.L. Electrochemical studies of iron(III) Schiff base complexes-II. Dimeric μ-oxo [FeIII(N2O2)]2O complexes. Polyhedron 1995, 14, 771–780. [Google Scholar] [CrossRef]
  24. Matosziuk, L.M.; Holbrook, R.J.; Manus, L.M.; Heffern, M.C.; Ratner, M.A.; Meade, T.J. Rational design of [Co(acacen)L2]+ inhibitors of protein function. J. Chem. Soc. Dalt. Trans. 2013, 42, 4002–4012. [Google Scholar] [CrossRef] [PubMed]
  25. Cini, R.; Cinquantini, A.; Orioli, P.L.; Mealli, C.; Sabat, M. The effect of d electron configuration on a mononuclear transition metal species of a Schiff-base ligand with a N2S2 donor set (sacacen). Can. J. Chem. 1984, 62, 2908–2913. [Google Scholar] [CrossRef]
  26. 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. [Google Scholar] [CrossRef]
  27. Ngo, K.T.; McKinnon, M.; Mahanti, B.; Narayanan, R.; Grills, D.C.; Ertem, M.Z.; Rochford, J. Turning on the Protonation-First Pathway for Electrocatalytic CO2 Reduction by Manganese Bipyridyl Tricarbonyl Complexes. J. Am. Chem. Soc. 2017, 139, 2604–2618. [Google Scholar] [CrossRef] [PubMed]
  28. 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. [Google Scholar] [CrossRef]
  29. Azcarate, I.; Costentin, C.; Robert, M.; Savéant, J.M. Through-Space Charge Interaction Substituent Effects in Molecular Catalysis Leading to the Design of the Most Efficient Catalyst of CO2-to-CO Electrochemical Conversion. J. Am. Chem. Soc. 2016, 138, 16639–16644. [Google Scholar] [CrossRef]
  30. Lam, Y.C.; Nielsen, R.J.; Gray, H.B.; Goddard, W.A. A Mn Bipyrimidine Catalyst Predicted to Reduce CO2 at Lower Overpotential. ACS Catal. 2015, 5, 2521–2528. [Google Scholar] [CrossRef] [Green Version]
  31. Queyriaux, N.; Abel, K.; Fize, J.; Pécaut, J.; Orio, M.; Hammarström, L. From non-innocent to guilty: On the role of redox-active ligands in the electro-assisted reduction of CO2 mediated by a cobalt(II)-polypyridyl complex. Sustain. Energy Fuels 2020, 4, 3668–3676. [Google Scholar] [CrossRef]
  32. Isse, A.A.; Gennaro, A.; Vianello, E.; Floriani, C. Electrochemical reduction of carbon dioxide catalyzed by [CoI(salophen)Li]. J. Mol. Catal. 1991, 70, 197–208. [Google Scholar] [CrossRef]
  33. Toniolo, D.; Scopelliti, R.; Zivkovic, I.; Mazzanti, M. Assembly of High-Spin [Fe3] Clusters by Ligand-Based Multielectron Reduction. J. Am. Chem. Soc. 2020, 142, 7301–7305. [Google Scholar] [CrossRef] [PubMed]
  34. Cometto, C.; Chen, L.; Mendoza, D.; Lassalle-Kaiser, B.; Lau, T.C.; Robert, M. An Iron Quaterpyridine Complex as Precursor for the Electrocatalytic Reduction of CO2 to Methane. ChemSusChem 2019, 12, 4500–4505. [Google Scholar] [CrossRef] [PubMed]
  35. 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. [Google Scholar] [CrossRef]
  36. 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. [Google Scholar] [CrossRef] [Green Version]
  37. Loewen, N.D.; Neelakantan, T.V.; Berben, L.A. Renewable Formate from C-H Bond Formation with CO2: Using Iron Carbonyl Clusters as Electrocatalysts. Acc. Chem. Res. 2017, 50, 2362–2370. [Google Scholar] [CrossRef] [PubMed]
  38. Solis, B.H.; Maher, A.G.; Honda, T.; Powers, D.C.; Nocera, D.G.; Hammes-Schiffer, S. Theoretical analysis of cobalt hangman porphyrins: Ligand dearomatization and mechanistic implications for hydrogen evolution. ACS Catal. 2014, 4, 4516–4526. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Iron complexes with tetradentate N2O2 ligands employed in this work. H2salophen: N,N′-bis(salicylaldehyde)-1,2-phenylenediamine; H2salen: N,N′-bis(salicylaldehyde)-1,2-ethylenediamine; H2acacen: N,N′-ethylene-bis(acetylacetoneimine); H2beacen: N,N′-ethylene-bis(benzoylacetoneimine); H2acacsalen: N,N′-ethylene-(acetylacetoneimine)(salicylideneimine).
Figure 1. Iron complexes with tetradentate N2O2 ligands employed in this work. H2salophen: N,N′-bis(salicylaldehyde)-1,2-phenylenediamine; H2salen: N,N′-bis(salicylaldehyde)-1,2-ethylenediamine; H2acacen: N,N′-ethylene-bis(acetylacetoneimine); H2beacen: N,N′-ethylene-bis(benzoylacetoneimine); H2acacsalen: N,N′-ethylene-(acetylacetoneimine)(salicylideneimine).
Energies 14 05723 g001
Figure 2. Cyclic voltammetries of 1 mM Fe(salen)Cl (left) and Fe(beacen)Cl (right) in acetonitrile (0.1 M Et4NBF4 electrolyte), under N2 (black, solid traces) and CO2 (dotted traces) atmosphere. The current of the voltammetries is normalized with respect to the peak current of the FeII→FeI reduction wave under dinitrogen, to visualize the enhancement in the presence of CO2. The glassy carbon (GC) disk working electrode, Pt counter electrode, and Ag/AgCl/NaCl (3 M) were the reference electrode; potentials were then converted to Fc+/Fc by internal calibration; scan rate 0.1 Vs−1.
Figure 2. Cyclic voltammetries of 1 mM Fe(salen)Cl (left) and Fe(beacen)Cl (right) in acetonitrile (0.1 M Et4NBF4 electrolyte), under N2 (black, solid traces) and CO2 (dotted traces) atmosphere. The current of the voltammetries is normalized with respect to the peak current of the FeII→FeI reduction wave under dinitrogen, to visualize the enhancement in the presence of CO2. The glassy carbon (GC) disk working electrode, Pt counter electrode, and Ag/AgCl/NaCl (3 M) were the reference electrode; potentials were then converted to Fc+/Fc by internal calibration; scan rate 0.1 Vs−1.
Energies 14 05723 g002
Figure 3. Redox equilibria of iron complexes. As previously observed for Fe (salophen)Cl [12], the loss of Cl should be considered relevant at the level of the FeI state under the conditions of the CV scans, in the absence of additional Cl sources.
Figure 3. Redox equilibria of iron complexes. As previously observed for Fe (salophen)Cl [12], the loss of Cl should be considered relevant at the level of the FeI state under the conditions of the CV scans, in the absence of additional Cl sources.
Energies 14 05723 g003
Figure 4. SEC-UV/Vis traces of Fe(salen)Cl (left) and of Fe(beacen)Cl (right) under N2 atmosphere (top panels) and under CO2 atmosphere (bottom panels), at the potentials corresponding to the generation of the formal FeI state (−2.21 and −2.24 V vs. Fc+/Fc for Fe(salen) and Fe(beacen), respectively).
Figure 4. SEC-UV/Vis traces of Fe(salen)Cl (left) and of Fe(beacen)Cl (right) under N2 atmosphere (top panels) and under CO2 atmosphere (bottom panels), at the potentials corresponding to the generation of the formal FeI state (−2.21 and −2.24 V vs. Fc+/Fc for Fe(salen) and Fe(beacen), respectively).
Energies 14 05723 g004
Figure 5. Cyclic voltammetries of 1 mM Fe(salen)Cl (left) and Fe(beacen)Cl (right) in acetonitrile (0.1 M Et4NBF4 electrolyte), under CO2 atmosphere, in the presence of 0.3 M trifluoroethanol (TFE, solid traces) or 0.5 M phenol (PhOH, dotted traces) as proton donors. The current of the voltammetries is normalized with respect to the peak current of the FeII→FeI reduction wave under dinitrogen in the absence of proton donors, in order to allow a direct comparison with Figure 2. The same experimental conditions as those reported in the caption of Figure 2. The inset on the right in Figure 5 shows the comparison of the voltametric traces with Fe(beacen)Cl in the presence of proton donors (blue traces) with the one registered in the absence of proton donors (black trace, see also Figure 2), that allows one to appreciate the current enhancement of the first catalytic process in the presence of proton donors; see previous discussion. In the presence of the proton donors, variation of the E1/2 of the FeIII/II couple and cathodic FeIII→FeII wave splitting are observed, see Figures S12 and S13 and Table S1. These observations are coherent with protonation equilibria of the complex, presumably at the ligand phenoxide and/or enolate oxygen sites, along with PhOH or TFE coordination to FeIII with competition with the Cl ligand for apical coordination see refs [11,12].
Figure 5. Cyclic voltammetries of 1 mM Fe(salen)Cl (left) and Fe(beacen)Cl (right) in acetonitrile (0.1 M Et4NBF4 electrolyte), under CO2 atmosphere, in the presence of 0.3 M trifluoroethanol (TFE, solid traces) or 0.5 M phenol (PhOH, dotted traces) as proton donors. The current of the voltammetries is normalized with respect to the peak current of the FeII→FeI reduction wave under dinitrogen in the absence of proton donors, in order to allow a direct comparison with Figure 2. The same experimental conditions as those reported in the caption of Figure 2. The inset on the right in Figure 5 shows the comparison of the voltametric traces with Fe(beacen)Cl in the presence of proton donors (blue traces) with the one registered in the absence of proton donors (black trace, see also Figure 2), that allows one to appreciate the current enhancement of the first catalytic process in the presence of proton donors; see previous discussion. In the presence of the proton donors, variation of the E1/2 of the FeIII/II couple and cathodic FeIII→FeII wave splitting are observed, see Figures S12 and S13 and Table S1. These observations are coherent with protonation equilibria of the complex, presumably at the ligand phenoxide and/or enolate oxygen sites, along with PhOH or TFE coordination to FeIII with competition with the Cl ligand for apical coordination see refs [11,12].
Energies 14 05723 g005
Figure 6. CO vs. H2 selectivity in CPE experiments with Iron complexes, depending on the proton donor. The selectivity was compared after the same charged passed, in particular at two electrons per iron catalyst (see Table 2), and used as performance indicators in the series.
Figure 6. CO vs. H2 selectivity in CPE experiments with Iron complexes, depending on the proton donor. The selectivity was compared after the same charged passed, in particular at two electrons per iron catalyst (see Table 2), and used as performance indicators in the series.
Energies 14 05723 g006
Figure 7. SEC-UV/Vis traces of Fe(salen)Cl under N2 atmosphere in the presence of 0.5 M PhOH (left) and 0.3 M TFE (right) proton donors, at the potentials corresponding to the generation of the formal FeI state (−2.21 V vs. Fc+/Fc).
Figure 7. SEC-UV/Vis traces of Fe(salen)Cl under N2 atmosphere in the presence of 0.5 M PhOH (left) and 0.3 M TFE (right) proton donors, at the potentials corresponding to the generation of the formal FeI state (−2.21 V vs. Fc+/Fc).
Energies 14 05723 g007
Table 1. Electrochemical and spectroscopic characterization of iron complexes employed in this work.
Table 1. Electrochemical and spectroscopic characterization of iron complexes employed in this work.
Iron ComplexE1/2 (V) vs. Fc+/Fc, V
(ΔE, mV)
λMAX FeI [a]−iCO2/iN2 [b]
FeIII/IIFeII/I
Fe(salophen)Cl [12]−0.69 (84)−2.00 [c]405−2.1
Fe(salen)Cl−0.77 (98)−2.21 (89)354−4.2
Fe(acacen)Cl−0.93 (93)−2.50 (68)288−5.6
Fe(beacen)Cl−0.83 (81)−2.24 (93)344−5.3
Fe(acacsalen)Cl−0.83 (91)−2.25 (131)308, 379−4.1
[a] From SEC-UV/Vis analysis. [b] Determined from the ratio of the peak current in the scan with CO2 and the peak current of the FeII/I couple under dinitrogen; while being a useful parameter to evaluate the catalytic activity, the –iCO2/iN2 ratio is dependent on the scan rate; values in this table refer to scan rate = 0.1 V s−1. [c] Quasi-reversible wave, and anodic backward peak only detectable at high scan rates (16 Vs−1).
Table 2. Electrochemical reduction of CO2 with iron complexes in the presence of phenol (PhOH) or trifluoroethanol (TFE) proton donors.
Table 2. Electrochemical reduction of CO2 with iron complexes in the presence of phenol (PhOH) or trifluoroethanol (TFE) proton donors.
Iron ComplexProton DonorEp (V vs. Fc+/Fc)−iCO2/iN2 [a]FY, CO
(FY, H2) [b]
CO vs. H2 Selectivity [b,c]
Fe(salophen)Cl [12]PhOH 0.5 M [11]−1.99−4.350 (1)98
Fe(salen)ClPhOH 0.5 M−2.25−7.640 (<0.5)>98 [d]
TFE 0.3 M−2.29−5.021 (<0.1)>99 [e]
Fe(acacen)ClPhOH 0.5 M−2.31−9.7 [f]5.5 (13.7)28
TFE 0.3 M−2.31−10.013.5 (0.55)96
Fe(beacen)ClPhOH 0.5 M−2.18−10.010 (14.4)41
TFE 0.3 M−2.15−7.842 (<0.1)>99
Fe(acacsalen)ClPhOH 0.5 M−2.19−7.4<1 (14.8)<1
TFE 0.3 M−2.31−5.818 (6.6)73
[a] Determined from CV traces; iCO2 is the cathodic peak current in the presence of CO2 and of the proton donor; iN2 is the cathodic peak current of the FeII/I couple under dinitrogen. [b] Determined after two electrons passed per iron center. [c] CO vs. H2 selectivity, defined as mol(CO)/{mol(CO)+mol(H2)}, and determined after two electrons passed per Iron center. [d] A selectivity >98% was maintained up to 25 electrons passed per iron center, with 4 turnovers for CO production. [e] A selectivity >99% was maintained up to 20 electrons passed per iron center, with 2.6 turnovers for CO production. [f] Ill-defined peak, due to the underlying current associated with the discharge of PhOH reduction.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bonetto, R.; Civettini, D.; Crisanti, F.; Sartorel, A. FeI Intermediates in N2O2 Schiff Base Complexes: Effect of Electronic Character of the Ligand and of the Proton Donor on the Reactivity with Carbon Dioxide. Energies 2021, 14, 5723. https://doi.org/10.3390/en14185723

AMA Style

Bonetto R, Civettini D, Crisanti F, Sartorel A. FeI Intermediates in N2O2 Schiff Base Complexes: Effect of Electronic Character of the Ligand and of the Proton Donor on the Reactivity with Carbon Dioxide. Energies. 2021; 14(18):5723. https://doi.org/10.3390/en14185723

Chicago/Turabian Style

Bonetto, Ruggero, Daniel Civettini, Francesco Crisanti, and Andrea Sartorel. 2021. "FeI Intermediates in N2O2 Schiff Base Complexes: Effect of Electronic Character of the Ligand and of the Proton Donor on the Reactivity with Carbon Dioxide" Energies 14, no. 18: 5723. https://doi.org/10.3390/en14185723

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

Article Metrics

Back to TopTop