Next Article in Journal
Probing the Bioinorganic Chemistry of Cu(I) with 111Ag Perturbed Angular Correlation (PAC) Spectroscopy
Next Article in Special Issue
Mono-Alkyl-Substituted Phosphinoboranes (HRP–BH2–NMe3) as Precursors for Poly(alkylphosphinoborane)s: Improved Synthesis and Comparative Study
Previous Article in Journal
The Analytical Application of Quenching Phenomena of CdTe Quantum Dot Nanoparticles
Previous Article in Special Issue
Carbon Quantum Dots: The Role of Surface Functional Groups and Proposed Mechanisms for Metal Ion Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Manganese(I) Diamine Electrocatalysts: Electrochemical Carbon Dioxide Reduction to Carbon Monoxide

1
Department of Chemistry, Oakland University, Rochester, MI 48309, USA
2
Department of Chemistry, Eckerd College, St. Petersburg, FL 33711, USA
3
Department of Biology, University of Tampa, Tampa, FL 33606, USA
4
School of Chemistry, University of Edinburgh, Edinburgh EH9 3FJ, UK
*
Author to whom correspondence should be addressed.
Inorganics 2023, 11(9), 374; https://doi.org/10.3390/inorganics11090374
Submission received: 1 September 2023 / Revised: 18 September 2023 / Accepted: 20 September 2023 / Published: 21 September 2023
(This article belongs to the Special Issue 10th Anniversary of Inorganics: Organometallic Chemistry)

Abstract

:
Novel organometallic complexes Mn(benzene-1,2-diamine)(CO)3Br, Mn-1, Mn(3-methylbenzene-1,2-diamine)(CO)3Br, Mn-2, and Re(benzene-1,2-diamine)(CO)3Cl, Re-1, have been synthesized and characterized by IR, UV/Vis, 1H-NMR, EA and HRMS. The structures of Mn-2 and Re-1 were confirmed by X-ray crystallography. The three novel compounds were studied for their electrocatalytic reduction of carbon dioxide to carbon monoxide using cyclic voltammetry in acetonitrile solutions. Controlled potential electrolysis was used to obtain information on faradaic yield, with product formation being confirmed by GC. Using earth-abundant manganese, compounds Mn-1 and Mn-2 display turnover frequencies of 108 s−1 and 82 s−1, respectively, amid selective production of carbon monoxide (faradaic yields ~85%), with minimal co-production of dihydrogen (<2%), and low overpotential of 0.18 V. The rhenium congener, Re-1, displays no activity as an electrocatalyst for carbon dioxide reduction under identical conditions.

Graphical Abstract

1. Introduction

Activation and utilization of carbon dioxide [1], a major greenhouse gas, has the potential to address two key contemporary environment related problems: global warming due to increasing CO2 emission and provision of carbon-neutral energy [2]. Electrocatalytic conversion of carbon dioxide to carbon monoxide [3] is considered one of the viable options for CO2 fixation [4], where liquid fuels [5] and other commodity chemicals can be made from the resulting carbon monoxide in combination with hydrogen gas using the Fischer–Tropsch technique [6].
Molecular organometallic electrocatalysts [7] for electroreduction of carbon dioxide [8] have been extensively reviewed [9], and trends within d-block metals have been elaborated [10,11,12]. Considering the need for ultimate scale-up of the catalytic process for CO2 reduction, designing catalysts using earth abundant transition metals [13], such as Fe [14,15], Ni [16,17], and Co [18,19], has received significant interest. Manganese is one of the most Earth abundant transition metals [20], and has been shown to replace rhenium [21] in CO2 reduction catalysts [22,23]. Hence, a wide-range of ligands have been attached to a manganese tricarbonyl halide core [24,25,26], including N-heterocyclic carbene [27,28], non-aromatic diamine [29], pincer ligands [30], and P-coordinating ligands [31]. There are a wide range of recent examples exploring structure-activity and mechanistic questions of these organomanganese electrocatalysts [32,33,34,35,36,37,38,39].
Thus, we began a study of aromatic diamine ligands, attached to either a manganese or rhenium tricarbonyl halide core. We report the novel compounds Mn(benzene-1,2-diamine)(CO)3Br, Mn-1, Mn(3-methylbenzene-1,2-diamine)(CO)3Br, Mn-2, and Re(benzene-1,2-diamine)(CO)3Cl, Re-1 (Figure 1). The historical compounds Mn(benzene-1,2-diamine)(CO)4 [40] and Re(o-benzoquinone diimine)(CO)3Cl [41] have each been reported on a single occasion, without any analysis for carbon dioxide electroreduction. In addition to synthesis and characterization, X-ray crystallography (Mn-2 and Re-1), electrochemistry, and electrocatalytic studies for the reduction of CO2 to CO by three novel organometallic diamine complexes are presented. The use of ligands bound to a metal via an amine group in successful organometallic electrocatalysts for carbon dioxide reduction is reported herein [42].
Electrochemistry will be utilized to determine turnover frequency (TOF) and overpotential from cyclic voltammetry (CV) of solutions of the compounds under an atmosphere of CO2.
T O F = 0.1992 F ν R T n p 3 n c a t 2 i c a t i p 2
CO2 (g) + 2HA (CH3CN) + 2e ⇋ CO (g) + H2O (CH3CN) + 2A (CH3CN)
= −0.12 (−0.0592 pKa) V vs. Fc+/Fc
where icat is the peak current of the catalyzed reaction and ncat is the number of electrons involved in that catalytic process, while ip is the peak current of the compound under non-catalytic conditions and np is the number of electrons involved in that process. F is the Faraday constant, ν is the applied scan rate, R is the universal gas constant, and T is the temperature in Kelvin. Subsequently, TOFs may be determined using Equation (1) [43].
Overpotential is also a useful metric for analysis of electrocatalysts, indicating how much additional potential is required beyond the thermodynamic amount for a given reaction. As we are exploring the reduction of carbon dioxide to carbon monoxide in the presence of a proton source, we will use Equation (2), where the thermodynamic potential is known to change with the strength of the acid chosen as the proton source [44].

2. Results and Discussion

2.1. Synthesis

Complexes Mn-1 and Mn-2 were synthesized by combining manganese pentacarbonyl bromide and the appropriate benzenediamine ligand in THF at reflux under an inert atmosphere. Complex Re-1 was synthesized by combining rhenium pentacarbonyl chloride and benzene-1,2-diamine in toluene at reflux under an inert atmosphere. Metal pentacarbonyl halide to ligand ratios were 1:1 to avoid the formation of complex salts, such as [Re(benzene-1,2-diamine)2(CO)3]Br [45]. Upon removal of THF/toluene by rotary evaporation, resolvation in minimal acetone was followed by crystallization from pentane, yielding dark green manganese compounds and a pale purple rhenium product.

2.2. Solid State Structure

Single crystals suitable for X-ray diffraction studies were grown via hexane vapor diffusion into a dichloromethane solution of Mn-2, and via slow evaporation of a diethyl ether solution of Re-1. Structures displayed tricarbonyl fac orientation (Figure 2 and Figure 3), with an average metal–carbon bond length of 1.79 Å for Mn-2 and 1.91 Å for Re-1, and metal–halide bond lengths of 2.55 Å and 2.51 Å, respectively.

2.3. Spectroscopy

UV/Visible absorption spectroscopy reveals similar MLCT bands for Mn-1 and Mn-2, with a molar attenuation coefficient ~20% greater for Mn-2 (Figure 4). Studies of tricarbonyl halide diimine compounds saw blue-shifts of 30–60 nm for replacement of Mn with Re [46]. Specifically with phenanthroline ligands, blue-shifts of ~45 nm are concomitant with a cathodic shift in reduction potential of ~250 mV [47]. As Re is switched to Mn herein, a cathodic shift in reduction potential of 450 mV is observed (Table 1), allowing a possible assignment of an MLCT to the absorption peak at 261 nm, a 98 nm blue shift (although IL π to π* absorptions are common in this region). This Re-1 putative MLCT has a molar attenuation coefficient of ~3-fold that of the manganese compounds, and is similar in magnitude to Re tricarbonyl halide diimines (~5000 M−1 cm−1) [48]. A weak absorption feature (88 M−1 cm−1) is seen for Re-1, occurring at ~180 nm red-shift compared to the MLCT of the manganese compounds (Figure 4, inset), which is likely a low probability MLCT [49], occurring at a wavelength very close to that seen for a much stronger MLCT seen in Re(o-benzoquinone diimine)(CO)3Cl [50].
IR spectroscopy of the carbonyl region displays three absorption peaks for all three compounds studied, displaying near identical wavenumbers for Mn-1 and Mn-2, correlating well to near identical reduction and oxidation potentials. Re-1 has a mean ṽCO some 16 cm−1 lower than the manganese compounds, indicative of stronger carbonyl binding due to greater electron density upon the central metal (supported by the observation of a more cathodic reduction). IR spectroscopy also displays the expected N-H bond absorbances in the 3100–3300 cm−1 region (Figures S9–S11), further supporting structural assignment.
NMR spectroscopy in acetone-d6 also confirms structural assignment, most clearly in the large downfield shift and loss of symmetry of the amine protons, seen in a single broad peak at 3.98 ppm in free benzene-1,2-diamine. Upon attachment of the ligand to the metal, the amine protons are seen as two equal integration irregular peaks at 6.58 and 4.88 ppm in Mn-1, and 6.88 and 5.50 ppm in Re-1 (Figures S12–S15). Attachment of 3-methylbenzene-1,2-diamine ligand (broad doublet at 3.87 ppm in free ligand) sees a similar shift upon attachment to manganese, yielding four equal integration irregular doublets at 6.58, 6.38, 4.88, and 4.67 ppm for Mn-2 with approximate splitting constants of 5 to 10 Hz (Figure S14). 13C NMR of Mn-2 (Figure S15) displays the anticipated 6 unique aromatic carbon atoms in the 140 to 127 ppm range, and a methyl carbon at 16.9 ppm (typically weak carbonyl carbon signals are obscured by the solvent).
The integration for the amine protons in Mn-1 and Re-1 is ~25% lower than anticipated, a feature that is preliminarily ascribed to exchangeability of these protons (the four amine proton peaks in the asymmetrical Mn-2 have only ~5% lower integration than expected). While no additional peaks are visible, concerns have been raised that this observation indicates that the benzene-1,2-diamine ligand is attached in a diimine manner. HRMS displayed two major peaks of similar size, one corresponding to the loss of halide and addition of an acetonitrile molecule, the second corresponding to just loss of halide [51]. No peaks were observed corresponding to compounds containing an o-benzoquinone diimine. Rhenium was confirmed by HRMS as present in Re-1 due to the observation of the predicted peak ratio due to the two primary rhenium isotopes. Elemental analysis also confirms the diamine molecular assignment of the three compounds studied herein. Table S1 details how alternative o-benzoquinone diimine compounds would not match the elemental analysis results for our compounds, lacking a match to the experimentally obtained values for hydrogen percentage. Additional evidence for diamine (rather diimine) ligands is seen in the bond lengths observed in the Xray crystallographic structures of Mn-2 and Re-1, where C-N bond lengths are 1.45 to 1.47 Å, matching the 1.47 Å average seen in amines, and far from the 1.28 Å average seen in imines [52].

2.4. Photochemistry

Organomanganese carbonyl compounds have long been reported as unstable upon light exposure [53], such that we carried out all our reactions and analyses with our compound in foil-wrapped glassware. Acetonitrile solutions of our three compounds were tracked for decomposition by UV/Vis over 90 min of exposure to ambient lab lighting. While Re-1 was unchanged, Mn-1 had a small amount degradation, ~3% change at λmax of 359 nm, and a small feature grew in at ~770 nm beyond 60 min of exposure. Mn-2 was more prone to decomposition, with ~5% change at λmax of 355 nm, and a degree of broadening of the main absorption feature (Figure S17).
Mn-1 and Mn-2 both display fluorescence upon excitation at their λmax wavelengths, 359 and 355 nm, respectively (Figure S18). Mn-1 emission is centered at 522 nm, while Mn-2 emission is centered at 541 nm, indicating greater energy loss to relaxation prior to emission, perhaps due to the lower symmetry molecule having more vibrational modes. Both manganese compounds had a minor (~20% the intensity of the main peak) secondary peak at longer wavelengths of 715 and 710 nm, respectively. Re-1 did not display any fluorescence upon excitation at either 261 or 543 nm.

2.5. Electrochemistry

All complexes exhibit a single irreversible reduction peak during cyclic voltammetry (CV), occurring at −2.30 V vs. Fc+/Fc under an argon atmosphere for Mn-1 (Figure S1), with the reduction potential of Mn-2 being slightly (50 mV) more negative than Mn-1 (Figure S2), indicative of a very modest increase in electron density, due to methylation upon the benzene ring [54] of the diamine. Support for this assignment can be found in the carbonyl group IR absorbance, where wavenumbers vary by just 2 cm−1 for the most labile CO. The reduction of Re-1 occurs a further 300 mV more negative, indicating more challenging reduction, and suggestive of more electron density upon the Re atom (as also indicated by the IR spectra of the carbonyl groups). Reductions are irreversible at the scan rates studied, appearing to be two-electron processes (based upon exhaustive controlled potential electrolysis of the compounds) and are assigned to loss of the halide ligand and formation of an anion [55].
The oxidation potential of the manganese compounds are almost identical at ~0.6 V vs. Fc+/Fc, while the oxidation of Re-1 occurs almost 300 mV more positive (Figure S3). All oxidations are irreversible and show large currents (greater than one or two electron processes) indicative of compound decomposition.

2.6. Electrocatalysis

All complexes were studied in acetonitrile under an atmosphere of CO2 in the presence of a proton source, trifluoroethanol (TFE). Addition of TFE to a solution of Mn-1 yielded enhancement in current at ca. −2.5 V (Figure 5), broadly similar to the effect seen with Mn-2 (Figure S4). The proposed catalytic mechanism (Scheme S1) suggests that initial reduction yields a catalyst that then cycles via a two electron two proton catalytic cycle, hence an enhancement of current is commonly interpreted as a catalytic process [8]. The reduction feature of Mn-1 (and Mn-2) occur at the same potential under a CO2 atmosphere; however, there is a change from a distinct peak under argon to a plateau under carbon dioxide, suggestive that the reduced complex binds substrate but no catalytic conversion occurs. This reduction feature is then seen to increase with increased concentration of TFE, an effect that is broadly absent in a similar study with Re-1 (Figure S5), and completely absent in the catalyst-free control experiment (Figure S6).
The addition of an external proton source to a solution of Mn-1 yields an increase in current due to the electrocatalytic reduction of carbon dioxide to carbon monoxide (vide infra). Homogeneous organometallic CO2 reduction electrocatalysts require either an internal proton source [56] or deliberate addition of an external proton source. The catalytically active species, obtained from electrochemical reduction, binds carbon dioxide but is unable to cleave one of the subsequent C-O bonds without the addition of a proton [57]. The catalytic current produced increases with increasing proton concentration (affected by either increased acid concentration or increased acid strength) [58]. A study of Mn-2 under a CO2 atmosphere with additions of water showed only a limited catalytic current (Figure S7); hence, our study focused upon the use of a stronger proton source, TFE. Additionally, a catalyst concentration study of Mn-2 in the presence of a fixed concentration of TFE shows a near linear response to increasing catalyst concentration, indicating a first-order process (Figure 6), and that catalysis is dependent upon the presence of the organomanganese electrocatalyst. The transfer of the working electrode used in a CO2/TFE study (with each manganese catalyst) into a catalyst-free solution containing CO2 and TFE displayed no catalytic process, supportive of assignment of homogeneous catalysis.
The icat/ip can be determined from CV data in the acid-independent region of catalysis [59], where icat and ip are the peak current of the catalyzed and non-catalyzed reactions, respectively. Entry into a true acid-independent region (by sufficient addition of acid to obtain a plateau in current) was not possible for our catalysts, due to solubility of the compounds becoming a problem at high values of added TFE. One solution to this problem would be to use less of a stronger (more acidic) proton source, although our concern preventing this was the competing reduction of the acid (or, rather, the carbonic acid that is made in situ) to hydrogen, due to our catalysts only being active at negative potentials where such competing reduction is an issue. However, the largest addition of TFE where no precipitation occurred (2.0 M) can be used as a proxy for the acid independent region by providing a reasonable minimum value for an icat/ip, and hence a minimum value for a TOF for our catalysts.
The icat/ip for Mn-1 was 8.3 and for Mn-2 it was 7.7 (Figure 7). These values may then be used in Equation (2) to find a turnover frequency (TOF) for the catalyst being studied [60], yielding values of 108 s−1 for Mn-1 and 82 s−1 for Mn-2. CV of Mn-1 vs. Mn-2 under catalytic conditions (in the presence of proton source and CO2) indicates that Mn-1 is the better catalyst, with an icat/ip value that is ~8% greater than for Mn-2. Given an electron donating methyl group on the ligand platform, Mn-2 was expected to produce a higher catalytic current than that of Mn-1 through a mild electronic effect that would enhance electron density upon the metal center and increase nucleophilic attacks upon the CO2 (Scheme S1) [61]. This lack of improvement suggests that the binding of CO2 by the catalyst is not rate-determining, but rather that the rate-determining step may be reductive cleavage of the Mn-CO (intermediate D to intermediate A, Scheme S1), a step that would be slower when metal center electron density is increased. With increased electron density on the metal, the Mn-CO bond would be stronger in Mn-2 due to greater back-bonding from metal to ligand, as compared to the case of Mn-1. The TOF values for our manganese compounds are somewhat low compared with the majority of published systems, with catalysis occurring at a disappointingly negative potential [62]. However, this initial example of diamine ligand use in carbon dioxide electrocatalyst design, as a proof of concept, could lead to exploration of the use of similar ligand platforms.
Our use of a weak acid proton source, namely trifluoroethanol (TFE) allows us to use Equation (2), with an estimated pKa of 35.4 for TFE [63], to arrive at a thermodynamic potential of −2.22 V. The potential at half peak height of the catalytic peak is used as a proxy for the E1/2 in irreversible systems (which the catalytic peak represents); in our case, a value ca. −2.4 V is observed, which represents a relatively low overpotential of ca. 0.18 V.
The limited electrocatalysis observed with Re-1 is consistent with examples where Re catalysts have TOF values around an order of magnitude lower than the analogous Mn compound [64]. In particular, in our case, Re-1 appears to reduce at too negative a potential, such that should any Re-CO adduct be forming, its reduction potential lies too negative (at/beyond the solvent window). Given that there is only a small increase in current at the edge of our solvent window (Figure S5), it is not possible to obtain an accurate icat/ip for Re-1, although it is likely <1.5.
CV of solutions of each catalyst in acetonitrile under argon with 2.00 M TFE displayed no peak, indicating that there is no competing proton reduction directly of the TFE [65]. Additionally, a CV of an acetonitrile electrolyte solution containing the same increments of TFE additions and saturated with carbon dioxide displays only a minor response (Figure S6), indicative of a negligible direct electrode response [66].
Controlled Potential Electrolysis (CPE) was performed upon all three compounds, as well as upon “catalyst-free” control experiment conditions, with the electrocatalysis products being quantified by headspace Gas Chromatography (GC) [67]. Comparison to authentic samples of CO and H2 (Figure S19) allowed the confirmation and quantification of the formation of carbon monoxide under these conditions with both manganese compounds. The performance of Mn-1 and Mn-2 are equivalent after electrolysis for 90 min, displaying faradaic yields of ~85% for CO production, with minimal co-production of H2 (Table 2). An effective turnover of ~7.5 for the manganese catalysts, where turnover is moles product divided by moles catalyst, establishes true catalysis with these manganese compounds, and given our 90 min CPE, equates to a turnover number (TON) of 5 h−1. Re-1 is inactive under the study conditions, producing ~2% the CO of the manganese compounds. CPE confirms the lack of activity of Re-1 that is implied by our earlier CV analysis.
CPE was stopped after 90 min for convenience and a desire to not cause pressure leaks in our reaction vessel during electrolysis. The organomanganese compounds were still active at this 90 min time point. Currents drop to around 80% of their initial value within a few minutes of initiation of electrolysis, and are from then mostly stable, slowly dropping to around 70–75% by the 90 min mark. If catalysis had ceased, then current would drop to zero, as it almost does for Re-1 and catalyst-free CPE. CV of the electrolysis solution upon completion of CPE still retains the catalytic peak, albeit somewhat diminished, to around 70% of the initial current value, indicative of some loss of organomanganese catalyst. An example FTIR spectra of a post electrolysis solution of Mn-2 clearly displays a tricarbonyl pattern (Figure S11). The pattern has peaks that are 1–2 cm−1 different from solid Mn-2, which could be an artefact of instrument precision or solvation, but does not preclude the formation of Mn(3-methylbenzene-1,2-diamine)(CO)3(CH3CN) [68].

3. Experiment/Characterization

Materials: All the reagents were used as received from the following: tetrabutylammonium hexafluorophosphate (Bu4APF6, Sigma, St. Louis, MI, USA), anhydrous acetonitrile (Sigma), manganese pentacarbonyl bromide (Mn(CO)5Br, Strem), rhenium pentacarbonyl chloride (Re(CO)5Cl, Sigma), benzene-1,2-diamine (TCI), 3-methylbenzene-1,2-diamine (Acros Organics).
Synthesis: All reactions were carried out in an inert atmosphere of argon using standard Schlenk techniques. All reactions involving manganese complexes were protected from light during reaction, storage and electrochemistry (using aluminum foil wrap).
Mn(benzene-1,2-diamine)(CO)3Br (Mn-1): Mn(CO)5Br (0.2020 g, 0.733 mmol) and 1,2-phenylenediamine (0.0800 g, 0.740 mmol) were combined in tetrahydrofuran (30 mL), and the mixture was purged in argon for 10 min before being stirred at reflux for one hour. A color change occurred during reflux, changing from the faint orange to black/green. The tetrahydrofuran was removed by use of a rotary evaporator until dryness, and acetone (2 mL) and pentane (50 mL) were added, forming a green precipitate. The product was cooled in an isotemperature freezer (−20 °C) and then filtered and rinsed with pentane. The final green solid was dried under reduced pressure for approximately 24 h, obtaining an 76% yield. IR (solid, CO): 2028 cm−1, 1931 cm−1, 1890 cm−1. MS (m/z): calc: 246.9915, found: 246.9913 (M-Br)+, calc: 288.0181, found: 288.0176 (M-Br+MeCN)+. UV/Vis in MeCN (ϵ, M−1 cm−1): 359 nm (1580). 1H NMR (300 MHz, (CD3)2CO, ppm) δ: 7.49 (2H, s), 7.30 (2H, s), 6.58 (2H, m), 4.88 (2H, m). Anal. Calc. for C9H8BrN2O3Mn: C, 33.06; H, 2.47; N, 8.57. Anal. Found: C, 33.34; H, 2.65; N, 8.42.
Mn(3-methylbenzene-1,2-diamine)(CO)3Br (Mn-2): Mn(CO)5Br (0.2010 g, 0.730 mmol) and 2,3-diaminotoluene (0.0901 g, 0.738 mmol) were combined in tetrahydrofuran (30 mL), and the mixture was purged with argon for ten minutes before stirring at reflux for one hour. A color change occurred during reflux from pale orange to dark green. The tetrahydrofuran was removed by rotary evaporator until dryness, and acetone (2 mL) and pentane (50 mL) were added, forming a brown/green precipitate. The product was cooled in an isotemp freezer (−20 °C) and then filtered and rinsed with pentane. The final brown/green solid was dried under reduced pressure for approximately 24 h, obtaining an 86% yield. IR (solid, CO): 2030 cm−1, 1932 cm−1, 1888 cm−1. MS (m/z): calc: 261.0071, found: 261.0064 (M-Br)+, calc: 302.0337, found: 302.0334 (M-Br+MeCN)+. UV/Vis in MeCN ((ϵ, M−1 cm−1): 355 nm (1930). 1H NMR (300 MHz, (CD3)2CO, ppm) δ: 7.32 (1H, m), 7.18 (2H, m), 6.58 (1H, d, 5.1 Hz), 6.38 (1H, d, 10.5 Hz) 4.88 (1H, d, 9 Hz), 4.67 (1H, d, 9.3 Hz), 2.54 (3H, s). 13C NMR (75 MHz, (CD3)2CO, ppm) δ: 139.7, 138.2, 135.6, 128.4, 126.9, 124.2, 16.9 (carbonyl C obscured by solvent). Anal. Calc. for C10H10BrMnN2O3: C, 35.22; H, 2.96; N, 8.21. Anal. Found: C, 35.50; H, 2.94; N, 8.06. Crystals suitable for X-ray diffraction were obtained by vapor diffusion of hexane into a solution of the complex in dichloromethane at −20 °C, CCDC 1893661.
Re(benzene-1,2-diamine)(CO)3Cl (Re-1): Re(CO)5Cl (0.2002 g, 0.553 mmol) and 1,2-phenylenediamine (0.0607 g, 0.562 mmol) were combined in toluene (50 mL) and the mixture was purged with argon for 10 min before being stirred at reflux for two hours. A color change occurred during reflux, changing from uncolored to purple. The toluene was removed by use of a rotary evaporator until dryness, and acetone (2 mL) and pentane (50 mL) were added, forming a pale purple precipitate. The product was cooled in an isotemperature freezer (−20 °C), and then filtered and rinsed with pentane. The final solid was dried under reduced pressure for approximately 24 h, obtaining an 80% yield. IR (solid, CO): 2025 cm−1, 1915 cm−1, 1861 cm−1. UV/Vis in MeCN (ϵ, M−1 cm−1): 261 nm (4700), 543 nm (88). 1H NMR (400 MHz, (CD3)2CO, ppm) δ: 7.50 (2H. m), 7.31 (2H, m), 6.88 (2H, d), 5.50 (2H, d). MS (m/z): calc: 379.0093, found: 379.0106 (M-Cl)+, calc: 420.0358, found: 420.0365 (M-Cl+MeCN)+. Anal. Calc. for C9H8ClN2O3Re: C, 26.12; H, 1.95; N, 6.77. Anal. Found: C, 26.37; H, 1.88; N, 6.82. Crystals suitable for X-ray diffraction were obtained by evaporation of diethyl ether solution of the complex at −20 °C, CCDC 1893660.
Electrochemistry: All cyclic voltammograms (CVs) were obtained by using a CHI Model 600E Potentiostat 3-electrode cell with a glassy carbon 3 mm diameter working electrode, platinum wire counter electrode, and silver external reference electrode (10 mM AgNO3 in electrolyte solution). The electrolyte solution for all experiments was 0.10 M tetra-n-butylammonium hexafluorophosphate (Bu4NPF6) in anhydrous acetonitrile (Aldrich). The potentials (E) at the working electrode in all CV’s are reported with respect to the ferrocenium/ferrocene couple in electrolyte solution. The ferrocenium/ferrocene couple data was collected daily at the end of each experiment, confirming potential and that appropriate resistance compensation had been applied. All CVs reported are background corrected, i.e., the scan with only electrolyte present was subtracted from the raw data. All background scans confirmed the sufficient removal of O2 as seen (Figure S20) by the absence of a reduction peak ca. −1.2 V [69]. CVs were collected under the flow of ultra-high purity argon or bone-dry carbon dioxide gases. Controlled Potential Electrolysis (CPE) experiments were performed using a two-compartment cell: one compartment containing a carbon rod working electrode (effective surface area = 5.5 cm2) and an Ag/AgNO3 pseudo-reference electrode, the other compartment containing a carbon rod counter electrode (effective surface area = 8.0 cm2). Head-space gas was tested after 90 min of electrolysis.
Instrumentation: NMR spectra were recorded on a 300 or 400 MHz Bruker NMR spectrometer. Mass spectra were recorded on an Agilent 6500 Series Q-TOF LC/MS in acetonitrile solution. Electronic absorption spectra were recorded on a Shimadzu UV-2600 UV-visible spectrometer. FTIR data was obtained using a Perkin Elmer 100 spectrometer, at 1.0 cm−1 resolution. Elemental analysis was carried out by Atlantic Microlabs (Norcross, GA, USA).

4. Conclusions

Two new earth abundant metal-based catalysts, Mn(benzene-1,2-diamine)(CO)3Br (Mn-1) and Mn(3-methylbenzene-1,2-diamine)(CO)3Br (Mn-2) have been synthesized and characterized, which rapidly catalyze the reduction of carbon dioxide to carbon monoxide at low overpotential. The novel Re congener Re(benzene-1,2-diamine)(CO)3Cl (Re-1) has been synthesized and characterized, but lacks such electrocatalytic activity. The reduction of Re-1 occurs 0.40 V more cathodic than the manganese complexes, likely indicating that if Re-1 were to undergo a similar mechanism, a catalytic peak further cathodic than the one seen for the manganese compounds would occur. We would likely not be able to see such a process due to competing direct reduction and the cathodic limit of the solvent window. [23,70] Use of diamine ligands could lead to an expansion of ligand options for organomanganese carbon dioxide reduction electrocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics11090374/s1, References [71,72,73] cited in Supplementary Materials.

Author Contributions

Conceptualization, B.D. and G.A.N.F.; methodology, B.D., B.A.C. and A.S.P.; validation, J.G.S. and E.A.F.; formal analysis, G.A.N.F.; investigation, B.D., B.A.C., A.S.P., A.J.G. and G.A.N.F.; data curation, L.T.M.; writing—original draft preparation, B.D. and G.A.N.F.; writing—review and editing G.A.N.F.; supervision, G.A.N.F.; project administration, G.A.N.F.; funding acquisition, G.A.N.F. and G.S.N. collected and solved single crystal XRD data. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Florida Space Grant Consortium.

Data Availability Statement

The data presented in this study are available within the article and supplementary material.

Acknowledgments

The authors gratefully acknowledge funding from the Florida Space Grant Consortium.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Sick, V. Spiers memorial lecture: CO2 utilization: Why, why now, and how? Faraday Discuss. 2021, 230, 9–29. [Google Scholar] [CrossRef]
  2. Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, H.D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.; Hille, R.; Kenis, P.J.; et al. Frontiers, opportunities, and challenges in biochemical and chemical catalysis of CO2 fixation. Chem. Rev. 2013, 113, 6621–6658. [Google Scholar] [CrossRef] [PubMed]
  3. Lim, R.J.; Xie, M.; Sk, M.A.; Lee, J.; Fisher, A.; Wang, X.; Lim, K.H. A review on the electrochemical reduction of CO2 in fuel cells, metal electrodes and molecular catalysts. Catal. Today 2014, 233, 169–180. [Google Scholar] [CrossRef]
  4. DuBois, D.L. Development of molecular electrocatalysts for energy storage. Inorg. Chem. 2014, 53, 3935–3960. [Google Scholar] [CrossRef] [PubMed]
  5. 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. [Google Scholar] [CrossRef] [PubMed]
  6. Underwood, A. Industrial synthesis of hydrocarbons from hydrogen and carbon monoxide. Ind. Eng. Chem. 1940, 32, 449–454. [Google Scholar] [CrossRef]
  7. Kinzel, N.W.; Werle, C.; Leitner, W. Transition metal complexes as catalysts for the electroconversion of CO2: An organometallic perspective. Angew. Chem. Int. Ed. 2021, 60, 11628–11686. [Google Scholar] [CrossRef]
  8. Costentin, C.; Robert, M.; Saveant, J. Catalysis of the electrochemical reduction of carbon dioxide. Chem. Soc. Rev. 2013, 42, 2423–2436. [Google Scholar] [CrossRef]
  9. Francke, R.; Schille, B.; Roemelt, M. Homogeneously catalyzed electroreduction of carbon dioxide—Methods, mechanisms, and catalysts. Chem. Rev. 2018, 118, 4631–4701. [Google Scholar] [CrossRef]
  10. Changcheng, C.; Nichola, A.W.; Machan, C.W. A look at periodic trends in d-block molecular electrocatalysts for CO2 reduction. Dalton Trans. 2019, 48, 9454–9468. [Google Scholar] [CrossRef]
  11. Grice, K.A. Carbon dioxide reduction with homogenous early transition metal complexes: Opportunities and challenges for developing CO2 catalysis. Coord. Chem. Rev. 2017, 336, 78–95. [Google Scholar] [CrossRef]
  12. Dalle, K.E.; Warnan, J.; Leung, J.J.; Reuillard, B.; Karmel, I.S.E. Reisner Electro- and solar-driven fuel synthesis with first row transition metal complexes. Chem. Rev. 2019, 119, 2752–2875. [Google Scholar] [CrossRef] [PubMed]
  13. Fisher, B.J.; Eisenberg, R. Electrocatalytic reduction of carbon dioxide by using macrocycles of nickel and cobalt. J. Am. Chem. Soc. 1980, 102, 7361–7363. [Google Scholar] [CrossRef]
  14. Costentin, C.; Drouet, S.; Robert, M.; Saveant, J. A local proton source enhances CO2 electroreduction to CO by a molecular Fe catalyst. Science 2012, 338, 90–94. [Google Scholar] [CrossRef]
  15. Oberem, E.; Roesel, A.F.; Rosas-Hernaandez, A.; Kull, T.; Fischer, S.; Spannenberg, A.; Junge, H.; Beller, M.; Ludwig, R.; Roemelt, M.; et al. Mechanistic insights into the electrochemical reduction of CO2 catalyzed by iron cyclopentadienone complexes. Organometallics 2019, 38, 1236–1247. [Google Scholar] [CrossRef]
  16. Froehlich, J.D.; Kubiak, C.P. Homogeneous CO2 reduction by Ni(cyclam) at a glassy carbon electrode. Inorg. Chem. 2012, 51, 3932–3934. [Google Scholar] [CrossRef]
  17. Mash, B.L.; Raghavan, A.; Ren, T. NiII complexes of C-substituted cyclam as efficient catalysts for reduction of CO2 to CO. Eur. J. Inorg. Chem 2019, 2019, 2065–2070. [Google Scholar] [CrossRef]
  18. Lacy, D.C.; McCrory, C.C.L.; Peters, J.C. Studies of cobalt-mediated electrocatalytic CO2 reduction using a redox-active ligand. Inorg. Chem. 2014, 53, 4980–4988. [Google Scholar] [CrossRef]
  19. Chapovetsky, A.; Do, T.H.; Haiges, R.; Takase, M.K.; Marinescu, S.C. Proton-assisted reduction of CO2 by cobalt aminopyridine macrocycles. J. Am. Chem. Soc. 2016, 138, 5765–5768. [Google Scholar] [CrossRef]
  20. Valyaev, D.A.; Lavigne, G.; Lugan, N. Manganese organometallic compounds in homogeneous catalysis: Past, present, and prospects. Coord. Chem. Rev. 2016, 308, 191–235. [Google Scholar] [CrossRef]
  21. Hawecker, J.; Lehn, J.; Ziessel, R.R. Photochemical and electrochemical reduction of carbon dioxide to carbon monoxide mediated by (2,2′-Bipyridine)tricarbonylchlororhenium(I) and related complexes as homogeneous catalysts. Helv. Chim. Acta 1986, 69, 1990–2012. [Google Scholar] [CrossRef]
  22. Bourrez, M.; Molton, F.; Chardon-Noblat, S.; Deronzier, A. [Mn(bipyridyl)(CO)3Br]: An abundant metal carbonyl complex as efficient electrocatalyst for CO2 reduction. Angew. Chem. Int. Ed. 2011, 50, 9903–9906. [Google Scholar] [CrossRef] [PubMed]
  23. 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. [Google Scholar] [CrossRef] [PubMed]
  24. Grills, D.C.; Ertem, M.Z.; McKinnon, M.; Ngo, K.T.; Rochford, J. Mechanistic aspects of CO2 reduction catalysis with manganese-based molecular catalysts. Coord. Chem. Rev. 2018, 374, 173–217. [Google Scholar] [CrossRef]
  25. Sinopoli; Porte, N.T.L.; Martinez, J.F.; Wasielewski, M.R.; Sohail, M. Manganese carbonyl complexes for CO2 reduction. Coord. Chem. Rev. 2018, 365, 60–74. [Google Scholar] [CrossRef]
  26. Stanbury, M.; Compain, J.-D.; Chardon-Noblat, S. Electro and photoreduction of CO2 driven by manganese-carbonyl molecular catalysts. Coord. Chem. Rev. 2018, 361, 120–137. [Google Scholar] [CrossRef]
  27. Agarwal, J.; Shaw, T.W.; Stanton, C.J.; Majetich, G.F.; Bocarsly, A.B.; Schaefer, H.F. NHC-Containing manganese(I) electrocatalysts for the two-electron reduction of CO2. Angew. Chem. Int. Ed. 2014, 53, 5152–5155. [Google Scholar] [CrossRef]
  28. Friães, S.; Realista, S.; Mourão, H.; Royo, B. N-Heterocyclic and mesoionic carbenes of manganese and rhenium in catalysis. Eur. J. Inorg. Chem. 2022, 2022, e202100884. [Google Scholar] [CrossRef]
  29. Zeng, Q.; Tory, J.; Hartl, F. Electrocatalytic reduction of carbon dioxide with a manganese(I) tricarbonyl complex containing a nonaromatic alpha-diimine ligand. Organometallics 2014, 33, 5002–5008. [Google Scholar] [CrossRef]
  30. Petersen, H.A.; Myren, T.H.T.; Luca, O.R. Redox-active manganese pincers for electrocatalytic CO2 reduction. Inorganics 2020, 8, 62. [Google Scholar] [CrossRef]
  31. Rao, G.K.; Pell, W.; Korobkova, I.; Richeson, D. Electrocatalytic reduction of CO2 using Mn complexes with unconventional coordination environments. Chem. Commun. 2016, 52, 8010–8013. [Google Scholar] [CrossRef] [PubMed]
  32. Siritanaratkul, B.; Eagle, C.; Cowan, A.J. Manganese carbonyl complexes as selective electrocatalysts for CO2 reduction in water and organic solvents. Acc. Chem. Res. 2022, 55, 955–965. [Google Scholar] [CrossRef] [PubMed]
  33. Scherpf, T.; Carr, C.R.; Donnelly, L.J.; Durawski, Z.S.; Gelfand, B.S.; Piers, W.E. A mesoionic carbene–pyridine bidentate ligand that improves stability in electrocatalytic CO2 reduction by a molecular manganese catalyst. Inorg. Chem. 2022, 61, 13644–13656. [Google Scholar] [CrossRef] [PubMed]
  34. Hong, W.; Luthra, M.; Jakobsen, J.B.; Madsen, M.R.; Castro, A.C.; Hammershøj, H.C.D.; Pedersen, S.U.; Balcells, D.; Skrydstrup, T.; Daasbjerg, K.; et al. Exploring the parameters controlling product selectivity in electrochemical CO2 reduction in competition with hydrogen evolution employing manganese bipyridine complexes. ACS Catal. 2023, 13, 3109–3119. [Google Scholar] [CrossRef]
  35. Singh, K.K.; Gerke, C.S.; Saund, S.S.; Zito, A.M.; Siegler, M.A.; Thoi, V.S. CO2 activation with manganese tricarbonyl complexes by an H-atom responsive benzimidazole ligand. Eur. J. Inorg. Chem. 2023, e202300796. [Google Scholar] [CrossRef]
  36. Cohen, K.Y.; Evans, R.; Bocarsly, A.B. Extending the π-system in MnI diimine tricarbonyl complexes: Impacts on photochemistry, electrochemistry, and CO2 catalytic reduction activity. Eur. J. Inorg. Chem. 2023, e202300435. [Google Scholar] [CrossRef]
  37. Florian, J.; Cole, J.M. Analyzing structure–Activity variations for Mn–carbonyl complexes in the reduction of CO2 to CO. Inorg. Chem. 2023, 62, 318–335. [Google Scholar] [CrossRef]
  38. Fernandez, S.; Franco, F.; Belmonte, M.M.; Friães, S.; Royo, B.; Luis, J.M.; Lloret-Fillol, J. Decoding the CO2 reduction mechanism of a highly active organometallic manganese electrocatlysts: Direct observation of a hydride intermediate and its implications. ACS Catal. 2023, 13, 10375–10385. [Google Scholar] [CrossRef]
  39. Eagle, C.; Neri, G.; Piercy, V.L.; Younis, K.; Siritanaratkul, B.; Cowan, A.J. A manganese complex on a gas diffusion electrode for selective CO2 to CO reduction. Sustain. Energy Fuels 2023, 7, 2301–2307. [Google Scholar] [CrossRef]
  40. Hieber, W.; Beck, W.; Zietler, G. Neuere anschauungen über reaktionsweisen der metallcarbonyle, insbesondere des mangancarbonyls. Angew. Chem. 1961, 73, 364–368. [Google Scholar] [CrossRef]
  41. Knor, G.; Leirer, M.; Synthesis, A.V. characterization and spectroscopic properties of 1,2-diiminetricarbonylrhenium(I)chloride complexes with o-benzoquinone diimines as ligands. J. Organomet. Chem. 2000, 610, 16–19. [Google Scholar] [CrossRef]
  42. Gonell, S.; Miller, A.J.M. Carbon dioxide electroreduction catalyzed by organometallic complexes. Adv. Organomet. Chem. 2018, 70, 1–69. [Google Scholar] [CrossRef]
  43. McKinnon, M.; Belkina, V.; Ngo, K.T.; Ertem, M.Z.; Grills, D.C.J. Rochford An investigation of electrocatalytic CO2 reduction using a manganese tricarbonyl biquinoline complex. Front. Chem. 2019, 7, 628. [Google Scholar] [CrossRef]
  44. 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. [Google Scholar] [CrossRef]
  45. Gerber, T.I.A.; Betz, R.; Booysen, I.N.; Potgieter, K.; Mayer, P. Coordination of bidentate aniline derivatives to the fac-[Re(CO)3]+ core. Polyhedron 2011, 30, 1739–1745. [Google Scholar] [CrossRef]
  46. Kottelat, E.; Lucarini, F.; Crochet, A.; Ruggi, A.; Zobi, F. Correlation of MLCTs of group 7 fac-[M(CO)3]+ complexes (M = Mn, Re) with bipyridine, pyridinylpyrazine, azopyridine, and pyridin-2-ylmethanimine type ligands for rational photoCORM design. Eur. J. Inorg. Chem. 2019, 2019, 3758–3768. [Google Scholar] [CrossRef]
  47. Kurtz, D.A.; Dhakal, B.; Hulme, R.J.; Nichol, G.S.; Felton, G.A.N. Correlations between photophysical and electrochemical properties for a series of new Mn carbonyl complexes containing substituted phenanthroline ligands. Inorg. Chim Acta 2015, 427, 22–26. [Google Scholar] [CrossRef]
  48. Kurtz, D.A.; Dhakal, B.; Donovan, E.S.; Nichol, G.S.; Felton, G.A.N. Non-photochemical synthesis of Re (diimine)(CO)2(L)Cl (L = phosphine or phosphite) compounds. Inorg. Chem. Commun. 2015, 59, 80–83. [Google Scholar] [CrossRef]
  49. Kurtz, D.A.; Brereton, K.R.; Ruoff, K.P.; Tang, H.M.; Felton, G.A.N.; Miller, A.J.M.; Dempsey, J.L. Bathochromic shifts in rhenium carbonyl dyes induced through destabilization of occupied orbitals. Inorg. Chem. 2018, 57, 5389–5399. [Google Scholar] [CrossRef]
  50. Kunkely, H.; Vogler, A. Absorption spectra and CT transitions of compounds containing ReI(CO)3Cl, S2+ and Se2+ as donors and o-benzoquinone diimine as acceptor. Inorg. Chem. Comm. 2007, 10, 1236–1238. [Google Scholar] [CrossRef]
  51. McIndoe, J.S.; Vikse, K.L. Assigning the ESI mass spectra of organometallic and coordination compounds. J. Mass Spectrom. 2019, 54, 466–479. [Google Scholar] [CrossRef] [PubMed]
  52. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. II 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  53. Kottelat, E.; Fabio, Z. Visible Light-activated PhotoCORMs. Inorganics 2017, 5, 24. [Google Scholar] [CrossRef]
  54. Donovan, E.S.; McCormick, J.J.; Nichol, G.S.; Felton, G.A.N. Cyclic voltammetric studies of chlorine-substituted benzenedithiolato-diiron hexacarbonyl electrocatalysts inspired by the [FeFe]-hydrogenase active site. Organometallics 2012, 31, 8067–8070. [Google Scholar] [CrossRef]
  55. Kurtz, D.A.; Dhakal, B.; McDonald, L.T.; Nichol, G.S.; Felton, G.A.N. Inter-ligand intramolecular through-space anisotropic shielding in a series of manganese carbonyl phosphorous compounds. Dalton Trans. 2019, 48, 14296–14935. [Google Scholar] [CrossRef] [PubMed]
  56. Franco, F.; Cometto, C.; Vallana, F.F.; Sordello, F.; Priola, E.; Minero, C.; Nervi, C.; Gobetto, R. A local proton source in a [Mn(bpy-R)(CO)3Br]-type redox catalyst enables CO2 reduction even in the absence of Bronsted acids. Chem. Commun. 2014, 50, 14670–14673. [Google Scholar] [CrossRef] [PubMed]
  57. Wong, K.; Chung, W.; Lau, C. The effect of weak Brønsted acids on the electrocatalytic reduction of carbon dioxide by a rhenium tricarbonyl bipyridyl complex. J. Electroanal. Chem. 1998, 453, 161–170. [Google Scholar] [CrossRef]
  58. Riplinger, C.; Carter, E.A. Influence of weak Brønsted acids on electrocatalytic CO2 reduction by manganese and rhenium bipyridine catalysts. ACS Catal. 2015, 5, 900–908. [Google Scholar] [CrossRef]
  59. Rountree, E.S.; McCarthy, B.D.; Eisenhart, T.T.; Dempsey, J.L. Evaluation of homogeneous electrocatalysts by cyclic voltammetry. Inorg. Chem. 2014, 53, 9983–10002. [Google Scholar] [CrossRef]
  60. Tezuka, M.; Iwasaki, M. Voltammetric Study on CO2 reduction electrocatalyzed by cobalt tetraphenylporphrine in DMF solution. Chem. Lett. 1993, 22, 427–430. [Google Scholar] [CrossRef]
  61. Bourrez, M.; Orio, M.; Molton, F.; Vezin, H.; Duboc, C.; Deronzier, A.; Chardon-Noblat, S. Pulsed-EPR evidence of a manganese(II) hydroxycarbonyl intermediate in the electrocatalytic reduction of carbon dioxide by a manganese bipyridyl derivative. Angew. Chem. Int. Ed. 2014, 53, 240–243. [Google Scholar] [CrossRef] [PubMed]
  62. Nie, W.; McCrory, C.C.L. Strategies for breaking molecular scaling relationships for the electrochemical CO2 reduction reaction. Dalton Trans. 2022, 51, 6993–7010. [Google Scholar] [CrossRef] [PubMed]
  63. Lam, Y.Y.; 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]
  64. Barrett, J.A.; Miller, C.J.; Kubiak, C.P. Electrochemical reduction of CO2 using group VII metal catalysts. Trends Chem. 2021, 3, 176–187. [Google Scholar] [CrossRef]
  65. Felton, G.A.N.; Glass, R.S.; Lichtenberger, D.L.; Evans, D.H. Iron-only hydrogenase mimics. Thermodynamic aspects of the use of electrochemistry to evaluate catalytic efficiency for hydrogen generation. Inorg. Chem. 2006, 45, 9181–9184. [Google Scholar] [CrossRef]
  66. McCarthy, B.D.; Martin, D.J.; Rountree, E.S.; Ullman, A.C.; Dempsey, J.L. Electrochemical reduction of Brønsted acids by glassy carbon in acetonitrile: Implications for electrocatalytic hydrogen evolution. Inorg. Chem. 2014, 53, 8350–8361. [Google Scholar] [CrossRef]
  67. Cook, T.D.; Tyler, S.F.; McGuire, C.M.; Zeller, M.; Fanwick, P.E.; Evans, D.H.; Peters, D.G.; Ren, T. Nickel complexes of C-substituted cyclams and their activity for CO2 and H+ reduction. ACS Omega 2017, 2, 3966–3976. [Google Scholar] [CrossRef]
  68. Cohen, K.Y.; Nedd, D.G.; Evans, R.; Bocarsly, A.B. Mechanistic insights into CO2 conversion to CO using cyano manganese complexes. Dalton Trans. 2023, 52, 7524–7537. [Google Scholar] [CrossRef]
  69. Donovan, E.S.; Felton, G.A.N. Electrochemical analysis of cyclopentadienylmetal carbonyl dimer complexes: Insight into the design of hydrogen-producing electrocatalysts. J. Organomet. Chem. 2011, 711, 25–34. [Google Scholar] [CrossRef]
  70. Mukherjee, J.; Siewert, I. Manganese and rhenium tricarbonyl complexes equipped with proton relays in the electrochemical CO2 reduction reaction. Eur. J. Inorg. Chem. 2020, e202000738. [Google Scholar] [CrossRef]
  71. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. Olex2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  72. Sheldrick, G.M. A short history of ShelX. Acta Cryst. 2008, A64, 339–341. [Google Scholar]
  73. Sheldrick, G.M. Crystal structure refinement with ShelXL. Acta Cryst. 2015, C27, 3–8. [Google Scholar]
Figure 1. Structures of the novel organometallic complexes studied herein: Mn(benzene-1,2-diamine)(CO)3Br, Mn-1, Mn(3-methylbenzene-1,2-diamine)(CO)3Br, Mn-2, and Re(benzene-1,2-diamine)(CO)3Cl, Re-1.
Figure 1. Structures of the novel organometallic complexes studied herein: Mn(benzene-1,2-diamine)(CO)3Br, Mn-1, Mn(3-methylbenzene-1,2-diamine)(CO)3Br, Mn-2, and Re(benzene-1,2-diamine)(CO)3Cl, Re-1.
Inorganics 11 00374 g001
Figure 2. ORTEP diagram of Mn-2 (30% probability thermal ellipsoids). The structure displays the presence of fac-[Mn(CO)3] with coordinated diamine and bromide ligand.
Figure 2. ORTEP diagram of Mn-2 (30% probability thermal ellipsoids). The structure displays the presence of fac-[Mn(CO)3] with coordinated diamine and bromide ligand.
Inorganics 11 00374 g002
Figure 3. ORTEP diagram of Re-1 (50% probability thermal ellipsoids). The structure displays the presence of fac-[Re(CO)3] with coordinated diamine and chloride ligand.
Figure 3. ORTEP diagram of Re-1 (50% probability thermal ellipsoids). The structure displays the presence of fac-[Re(CO)3] with coordinated diamine and chloride ligand.
Inorganics 11 00374 g003
Figure 4. UV/Vis spectra of Mn-1 (green), Mn-2 (brown) and Re-1 (purple), all 0.200 mM in acetonitrile. Inset: Re-1 concentration increased to 1.00 mM in acetonitrile.
Figure 4. UV/Vis spectra of Mn-1 (green), Mn-2 (brown) and Re-1 (purple), all 0.200 mM in acetonitrile. Inset: Re-1 concentration increased to 1.00 mM in acetonitrile.
Inorganics 11 00374 g004
Figure 5. Cyclic voltammetry of 1.010 mM of Mn-1 (blue) with subsequent additions of trifluoroethanol, TFE: 0.10 M, 0.20 M, 0.50 M, 1.00 M, 1.50 M, 2.00 M. Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Figure 5. Cyclic voltammetry of 1.010 mM of Mn-1 (blue) with subsequent additions of trifluoroethanol, TFE: 0.10 M, 0.20 M, 0.50 M, 1.00 M, 1.50 M, 2.00 M. Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Inorganics 11 00374 g005
Figure 6. Cyclic voltammetry of indicated concentrations of Mn-2 with 2.00 M trifluoroethanol. Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Figure 6. Cyclic voltammetry of indicated concentrations of Mn-2 with 2.00 M trifluoroethanol. Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Inorganics 11 00374 g006
Figure 7. Cyclic voltammetry of 2.00 M TFE (dotted line); 2.00 M TFE + 1.010 mM of Mn-1 (dark green); 2.00 M TFE + 1.002 mM of Mn-2 (blue); 2.00 M TFE + 0.992 mM of Re-1 (brown). Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Figure 7. Cyclic voltammetry of 2.00 M TFE (dotted line); 2.00 M TFE + 1.010 mM of Mn-1 (dark green); 2.00 M TFE + 1.002 mM of Mn-2 (blue); 2.00 M TFE + 0.992 mM of Re-1 (brown). Recorded in 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a 3.00 mm diameter glassy carbon electrode vs. Fc+/Fc at 0.100 V/s.
Inorganics 11 00374 g007
Table 1. Reduction/oxidation potentials, infrared and UV/Vis data for Mn-1, Mn-2, and Re-1. IR absorption peaks are attributed to CO stretching modes.
Table 1. Reduction/oxidation potentials, infrared and UV/Vis data for Mn-1, Mn-2, and Re-1. IR absorption peaks are attributed to CO stretching modes.
ComplexEpc1 (V) aEpa1 (V) aCO (cm−1) bMean CO
(cm−1)
MLCT
(nm) c
ϵ
M−1 cm−1
Mn-1−2.300.602028, 1931, 189019503591580
Mn-2−2.350.622030, 1932, 188819503551930
Re-1−2.750.892025, 1915, 18611934261, 5434700, 88
a Potential at peak of first reduction/oxidation wave vs. Fc/Fc+ at 0.100 V s−1 in 0.100 M n-Bu4NPF6 in anhydrous acetonitrile. b Recorded in solid state. c Recorded in acetonitrile.
Table 2. CPE data for Mn-1, Mn-2, and Re-1, compared to catalyst-free conditions a.
Table 2. CPE data for Mn-1, Mn-2, and Re-1, compared to catalyst-free conditions a.
Q (C)µmol COFE b CO (%)µmol H2FE b H2 (%)
Mn-135 ± 1.4157 ± 586 ± 32.9 ± 0.91.6 ± 0.8
Mn-233 ± 2148 ± 685 ± 42.6 ± 0.71.5 ± 0.7
Re-12.2 ± 0.63.1 ± 0.527 ± 52.4 ± 0.421 ± 4
Catalyst-free1.6 ± 0.40.6 ± 0.47 ± 45.4 ± 0.865 ± 10
a 0.500 mM (20.0 µmol) of catalyst in each case, in 1.50 M TFE, 0.100 M Bu4NPF6 in dry CO2-saturated acetonitrile on a carbon rod electrode at −2.20 V vs. Fc/Fc+ for 90 min (mean of three trials). b FE = Faradaic efficiency = charge to form product/total charge passed.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dhakal, B.; Corbin, B.A.; Sosa Parada, A.; Sakai, J.G.; Felton, E.A.; McDonald, L.T.; Gross, A.J.; Nichol, G.S.; Felton, G.A.N. Manganese(I) Diamine Electrocatalysts: Electrochemical Carbon Dioxide Reduction to Carbon Monoxide. Inorganics 2023, 11, 374. https://doi.org/10.3390/inorganics11090374

AMA Style

Dhakal B, Corbin BA, Sosa Parada A, Sakai JG, Felton EA, McDonald LT, Gross AJ, Nichol GS, Felton GAN. Manganese(I) Diamine Electrocatalysts: Electrochemical Carbon Dioxide Reduction to Carbon Monoxide. Inorganics. 2023; 11(9):374. https://doi.org/10.3390/inorganics11090374

Chicago/Turabian Style

Dhakal, Badrinath, Brooke A. Corbin, Alberto Sosa Parada, Jonathan G. Sakai, Emily A. Felton, Lauren T. McDonald, Anthony J. Gross, Gary S. Nichol, and Greg A. N. Felton. 2023. "Manganese(I) Diamine Electrocatalysts: Electrochemical Carbon Dioxide Reduction to Carbon Monoxide" Inorganics 11, no. 9: 374. https://doi.org/10.3390/inorganics11090374

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