Next Article in Journal
Experiments and Calculation on New N,N-bis-Tetrahydroacridines
Previous Article in Journal
Effect of Tourmaline Addition on the Anti-Poisoning Performance of MnCeOx@TiO2 Catalyst for Low-Temperature Selective Catalytic Reduction of NOx
Previous Article in Special Issue
Synthesis and Characterization of Transition Metal Complexes Supported by Phosphorus Ligands Obtained Using Hydrophosphination of Cyclic Internal Alkenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Electrochemical Study of Gold(I) Carbene Complexes

by
Andrea Rodríguez-Rubio
,
Álvaro Yuste
,
Tomás Torroba
,
Gabriel García-Herbosa
and
José V. Cuevas-Vicario
*
Departamento de Química, Facultad de Ciencias, Universidad de Burgos, 09001 Burgos, Spain
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(17), 4081; https://doi.org/10.3390/molecules29174081 (registering DOI)
Submission received: 20 June 2024 / Revised: 22 August 2024 / Accepted: 26 August 2024 / Published: 28 August 2024
(This article belongs to the Special Issue Advances in Coordination Chemistry 2.0)

Abstract

:
In this work, we have prepared and characterized some gold compounds wearing a N-heterocyclic carbene (NHC) ligand as well as alkynyl derivatives with different substituents. The study of their electrochemical behavior reveals that these complexes show an irreversible wave at potentials ranging between −2.79 and −2.91 V, referenced to the ferrocenium/ferrocene pair. DFT calculations indicate that the reduction occurs mainly on the aryl−C≡C fragment. The cyclic voltammetry experiments under CO2 atmosphere show an increase in the faradaic current of the reduction wave compared to the experiments under argon atmosphere, indicating a possible catalytic activity towards the carbon dioxide reduction reaction (CO2RR).

Graphical Abstract

1. Introduction

In recent years, gold(I) N-heterocyclic carbenes (NHCs) have emerged as interesting compounds with several applications [1,2]. Among these applications, several uses in bioinorganic applications have been recently reviewed [3,4,5]. Examples of these complexes behave as therapeutics for cancer treatment [6,7,8] or antibacterial agents [9]. In addition to these biological applications, gold(I) NHCs are involved as catalysts in a wide number of reactions [10,11]. In some recent works, the use of gold(I) NHCs as catalysts has been described for different reactions such as asymmetric catalysis [12,13,14], the synthesis of small rings [15], C–H bond activations [16,17], cross-coupling reactions [18], hydration reactions [19] and aqueous-phase catalysis [20], showing the high interest in gold complexes as catalysts for different reactions.
The carbon dioxide reduction reaction (CO2RR) is a very interesting possibility for transforming CO2 into valuable fuels and chemicals using renewable energy [21]. Several approaches such as biological [22,23,24,25], thermochemical [26,27,28], photochemical [29,30,31,32] and electrochemical [25,33,34,35,36,37,38,39,40,41] are under study to convert CO2 into formic acid, carbon monoxide, methanol, methane or other compounds with higher numbers of carbon atoms. Among these different approaches, electrochemical reduction has some advantages, such as the possibility of working at mild conditions of pressure and temperature, product selectivity by controlling the applied potential and the possibility of obtaining the necessary electricity from renewable sources [33,42]. The mild conditions used in the electrochemical reduction open up the possibility of using homogenous catalysis. Several examples of homogeneous electrocatalysis with transition metals have been reported [43,44,45].
In the context of the use of CO2 as a feedstock, several activation reactions of this molecule have been reported. In these reactions, gold(I) NHCs catalyze the carboxylation of activated C–H bonds [46,47], carboxylative cyclization [47,48,49] or the formation of a trigold carbonate complex from atmospheric CO2 [50]. This activation of the CO2 may encourage the search for gold complexes as possible catalysts of the CO2RR. Although gold is used in some instances as an electrode in this process, few examples involving organometallics or coordination compounds [51] or clusters [51,52,53,54,55,56,57,58] are reported. In all cases, the reaction product was CO, but a small amount of formic acid could be detected in the solution [53]. In the present work, we describe the synthesis and characterization of alkynyl-gold(I) complexes with N-heterocycle carbene ligands, and we explore their reactivity towards the electrocatalyzed CO2RR.

2. Results and Discussion

The reaction of a gold complex [Au(SIPr)Cl] with different arylalkynes in the presence of sodium acetate following the procedure described by Nolan et al. [59] yielded the compounds 15 (Scheme 1).
These complexes were characterized by FT-IR, 1H NMR spectroscopies and elemental analysis. The infrared spectra of these compounds feature a characteristic band at ~2115 cm−1 for ῦ(C≡C). The NMR spectra in DMSO-d6 (Figures S1–S4) show a characteristic sharp singlet at 4.11–4.07 ppm that corresponds to the equivalent four nuclei of the hydrogen atoms of the methylene groups of the NHC ring. The isopropyl signals appear as a septuplet at 3.17–3.10 ppm and two doublets at 1.43–1.35 and 1.42–1.31 ppm. The aromatic protons from the aryl-alkyne and the diisopropyl-phenyl rings appear mixed in the 8.55–6.91 ppm interval.

2.1. X-ray Structure Analyses

Single crystals suitable for X-ray diffraction (XRD) analysis were obtained by slow diffusion of hexane over a concentrated solution of 2 in CHCl3. The X-ray molecular structure of this complex and its atom numbering are shown in Figure 1. Table 1 collects the crystallographic data, and Table 2 gathers some structural features of these complexes. This compound crystallized in the orthorhombic system with the P212121 space group. The asymmetric unit contains two molecules. In one of these molecules, the ring of the NHC and the plane of the phenanthrene form an angle of 10.15°, and in the other molecule, the angle involving the same fragments is 63.79°. The molecular structure found is characterized by the gold atom being two-coordinated by two carbon atoms in an almost linear arrangement with C–Au–C angles of 176.1 (3) and 173.8 (3). The couple of fragments interact with each other through two C–H···H–C interactions [60]. One of these interactions involves two methyl groups of different fragments, and the other interaction is established between a C−H of an isopropyl group of one fragment and a C−H of the phenanthrenyl group of the other fragment (see Figure S5). The fragments of the asymmetric unit interact with other analogous complexes of gold to build up the crystal. These interactions are C−H···H−C and C−H···π towards the aromatic rings and the triple C≡C bond (see Figures S6 and S7).

2.2. Electrochemical Measurements

The electrochemistry of these complexes in acetonitrile solution was studied using cyclic voltammetry and Bu4NPF6 as a supporting electrolyte. The values of the cyclic voltammetry are collected in Table 3. In these measurements, a AgCl/Ag (3 M KCl) reference electrode was used, and at the end of the experiment, ferrocene was added as an internal calibrant. With the only exception being 1, upon scanning to negative potentials under argon, all complexes display irreversible reduction waves at negative values of potential under argon atmosphere (blue line in Figure 2 and Figures S8–S12). Compounds 24 display an irreversible reduction wave between −2.87 and −2.91 V versus the ferrocenium/ferrocene (Fc+/Fc) couple. Compound 5 has an irreversible reduction at −2.79 V, and complex 1 shows no similar reduction wave. The irreversible characteristic of the reduction wave is related to the evolution of the reduction product. These values are close to the other reported values for carbene complexes of gold(I) with amido ligands [61,62]. Gold(I) carbene complexes with halogen atoms completing the coordination of the metal (instead of the alkynyl ligands reported in this work) usually show less negative values for the reduction wave ranging from −1.22 V to −1.82 V [63,64,65], although there is reported an example in which the carbene is a ferrocenyl-substituted 1,2,3-triazolylidene ligand, showing a reduction wave at −2.47 V [66]. In gold complexes, this wave is generally assigned to the reduction of Au(I) to Au(0) [67], but we found (see below, quantum chemical calculations and Table S1) that the topology of the lowest unoccupied molecular orbital (LUMO) of these complexes locates this molecular orbital mainly over the C≡C–Aryl fragments (a participation ranging between 81.40 and 91.03% of the orbitals of the C≡C–Aryl on the LUMO) with little participation of the atomic orbitals of the gold (ranging between 7.41 and 15.65%). As the LUMO of these complexes is mainly spread over the aryl–C≡C fragment, the reduction process occurs mainly in these fragments of the complexes, although with a small participation of the gold center. In the case of complex 1, the high value of the energy of the LUMO can explain the lack of value for the reduction potential under our experimental conditions because this value should appear out of the window of measurement of the experiment [68,69,70,71].
Several NHC complexes of transition metals that are electrocatalysts of the CO2RR yielding CO have been reported. Re(I) with pyridyl–NHC or pyrimidyl−NHC chelating ligands are efficient catalysts in the reduction of CO2 to CO [72,73]. Related Mn(I) complexes with pyridyl–NHC ligands [74], and the same metal with chelating bis-NHC ligands [75,76], display electrocatalytic activity for the reduction of CO2 to CO. Fe(II) complexes with analogous pyridyl−NHC ligands or chelating bis-NHC and terpyridine as redox-active ligands also show electrocatalytic activity towards the CO2RR, yielding CO [77,78]. The combination of pyridyl–NHC and terpyridine ligands coordinated to Ru(II) generates electrocatalysts for the same reaction as the analogous combination of Fe(II) [79], but a similar complex immobilized on an N-doped porous carbon electrode reduced the CO2 to ethanol [80]. In addition to these complexes, bis-NHC ligands with different connectors such as pyridine (CNC pincer ligand) [81], xanthene (COC flexible and hemilabile ligand) [82] or methylene [83] show this electrocatalytic activity as well. Complexes of Co(II) [84] or Ni(II) [85] with macrocyclic ligands that were composed of a bipyridine fragment and two NHCs were active in this reaction. Pd(II) complexes with bis-NHC ligands bearing unique onium functionalities facilitate CO2 coordination to the catalytic center and enhance catalytic selectivity for the conversion of CO2 to CO compared to H2 production [86]. To the best of our knowledge, no gold NHC complexes with this activity have been reported. When the cyclic voltammetry experiments are performed under CO2 atmosphere, an increase in the cathodic current is observed (orange lines in Figure 2 and Figures S4−S8). This variation on the voltammogram is indicative of a CO2-electrocatalyzed reduction. The comparison of the electrochemical activity of the complexes can be analyzed from the values of the iCO2/iAr ratio. These values are ranging between 1.77 and 2.61. Figure 2 shows the results obtained for complex 2 as a representative example.

2.3. Quantum Chemical Calculations

DFT calculations were performed in order to obtain a better understanding of the electronic structure of these complexes (see Section 3 for details). Table 2 gathers some calculated structural values of the optimized structure for complex 2 in its ground state (S0), which showed only positive frequencies, and the comparison with the experimental XRD values, showing a good agreement between the structural parameters, validating the level of theory. Figure 3 shows the energy levels of the frontier molecular orbitals (FMOs) for all complexes. For complexes 15, both the highest occupied molecular orbital (HOMO) and the LUMO are mainly located on the aryl-alkynyl fragment, with some small participation of the metal center, as can be seen in Figure 4 and Figures S9–S12 and Table S1. This topology of the LUMO indicates that the entry of the electron in the reduction process occurs mainly on this fragment. The high value of the energy of the LUMO in complex 1 correlates well with the high value of the reduction potential for complex 1 [68,69,70,71].

3. Materials and Methods

Unless otherwise stated, all reagents were purchased from commercial suppliers and used without further purification. Specifically, 2-ethynylpyridine, 1-bromo-4-ethynylbenzene and 3-Chloro-1-ethynylbenzene were purchased from Thermo Scientific (Waltham, MA, USA), 9-ethynylphenanthrene was purchased from TCI and HAuCl4·3H2O was purchased from Johnson Matthey (London, UK). Complexes [Au(NHC)(Cl)] and 1 were prepared following previously reported procedures [59]. A JASCO FT/IR-4200 spectrometer equipped with an ATR accessory model JASCO ATR PRO ONE was used for the recording of infrared spectra. Nuclear magnetic resonance (NMR) spectra were recorded with Varian Mercury-300 and BRUKER AVANCE III HD 300 MHz spectrometers. These measurements were performed at room temperature (25 °C) using DMSO-d6 as a solvent. The reference used for reporting the chemical shifts (δ) was the residual solvent peaks, and the units were parts per million (ppm), rounded to the nearest 0.01. For signals with coupling between nuclei, spin–spin coupling constants (J) were given in Hz, rounded to the nearest 0.1 Hz, and peak multiplicity was indicated as s (singlet), d (doublet), t (triplet), q (quartet), sept (septuplet), m (multiplet) and br (broad). The purity of the complexes was tested through elemental analyses that were performed with a Thermo Fisher Scientific EA Flash 2000 elemental microanalyzer.

3.1. Electrochemical Measurements

Cyclic voltammetry measurements were performed using portable potentiostat/galvanostat PalmSens3 (PalmSens, Houten, The Netherlands) equipment controlled by the software PSTrace4 Version 4.4.2. These experiments of cyclic voltammetry were carried out under argon or CO2 atmosphere. The cell had three electrodes; one of them was a glassy carbon disc (diameter = 3 mm) acting as the working electrode, another was a platinum wire acting as the auxiliary electrode and the third electrode was a Ag/AgCl (MF-2052 BASi) reference electrode separated from the bulk solution by a Vycor frit. At the beginning of the experiments, oxygen was removed from the solution by bubbling argon for 10 min and keeping the current of argon constant throughout the whole experiment. The solvent used in the solutions was MeCN, and the concentration was 5 × 10−4 M in all cases. [n-Bu4N][PF6] (0.1 M) was used as the supporting electrolyte. The scan rate of the cyclic voltammetry was 100 mV·s–1 in a clockwise direction. At the end of all the experiments, ferrocene was added as the internal reference. The potential experimentally determined for the redox couple Fc+/Fc was E1/2° = 0.476 ± 0.002 V vs. Ag/AgCl.

3.2. Quantum Chemical Calculations

DFT calculations were performed using Becke’s three-parameter B3LYP exchange–correlation functional [87,88] implemented in Gaussian16 [89]. The basis sets used to define the atoms were LANL2DZ [90] for Au and 6-31G (d,p) [91,92] for the other atoms. The empirical dispersion correction was taken into account using Grimme’s dispersion with Becke–Johnson damping, GD3BJ [93,94]. The solvent (MeCN) effects were considered within the self-consistent reaction field theory using the solvation model SMD, as described by Marenich et al [95].

3.3. X-ray Crystallography

A single crystal of 2 was mounted on a glass fiber and transferred to a Bruker X8 APEX II CCD-based diffractometer equipped with a graphite-monochromated Cu Kα radiation source (λ = 0.71073 Å). The highly redundant data sets were integrated using SAINT [96] and corrected for Lorentz and polarization effects. The absorption correction was based on the function fitting to the empirical transmission surface, as sampled by multiple equivalent measurements with the program SADABS [97]. The software package SHELXTL (version 6.10) [98] was used for space group determination, structure solution and refinement with full-matrix least-squares methods based on F2. A successful solution by direct methods provided most non-hydrogen atoms from the E map. The remaining non-hydrogen atoms were located in an alternating series of least-squares cycles and difference Fourier maps. All non-hydrogen atoms were refined with anisotropic displacement coefficients. Hydrogen atoms were placed using a “riding model” and included in the refinement at calculated positions. CCDC 2363526 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/ (accessed on 25 August 2024) (or from the CCDC, Cambridge, UK).

3.4. Synthesis of the Complexes

Complexes 2, 3 and 5 [Au(SIPr)(C≡C–aryl)] were prepared following a general synthetic method [59] with a slight modification. Briefly, the starting material [Au(SIPr)Cl] (0.08 mmol), sodium acetate (0.24 mmol) and the corresponding alkyne (0.19 mmol) were deposited in a vial and suspended in EtOH (2 mL). The reaction mixture was stirred at room temperature for 16 h and protected from light. After that time, the solvent was removed under vacuum. The solid residue was dissolved in CH2Cl2 (2 mL), and the solution was filtered through Kieselgur. The filtrate was concentrated under vacuum, and the product was precipitated by adding n-hexane (3 mL). The solid was collected, washed three times with diethyl ether and dried under vacuum.
[Au(SIPr)(C≡C(2-phenanthrenyl)] (2). Yield 58%. Anal. Calcd for C43H47N2Au: C, 65.47; H, 6.02; N, 3.55. Found: C, 65.41; H, 6.27; N, 3.15. Ir (KBr, cm−1); 2962(νC–H), 2104(νC≡C), 1490(νCCarom), 1475(νCCarom). 1H NMR (300 MHz, DMSO-d6) δ 7.44 (dt, J = 7.4, 5.7 Hz, 2H), 7.34 (d, J = 7.7 Hz, 4H), 7.18–7.00 (m, 5H), 4.07 (s, 4H), 3.11 (sept, J = 6.8 Hz, 4H), 1.36 (d, J = 6.8 Hz, 12H), 1.32 (d, J = 6.9 Hz, 12H). 13C NMR (75 MHz, DMSO-d6) δ 208.47, 146.56, 146.09, 138.26, 134.51, 131.39, 131.34, 129.57, 129.46, 129.23, 128.55, 127.81, 126.86, 126.67, 124.34, 122.90, 122.61, 122.06, 101.31, 53.78, 28.26, 28.25, 25.01, 23.70.
[Au(SIPr)(C≡C(3-chlorophenyl] (3). Yield 32 %. Anal. Calcd. For C35H42N2AuCl: C, 58.13; H, 5.87; N, 3.87. Found: C, 58.14; H, 5.90; N, 3.69. Ir (KBr, cm−1): 2960(νC–H), 2114(νC≡C), 1588(νCCarom), 1491(νCCarom), 1275(νC–N), 684(νC–Cl). 1H NMR (300 MHz, DMSO-d6) δ 7.46 (dd, J = 8.5, 6.9 Hz, 2H), 7.34 (d, J = 7.7 Hz, 4H), 7.16–7.12 (m, 2H), 7.07 (d, J = 1.5 Hz, 1H), 7.04–6.99 (m, 1H), 4.08 (s, 4H), 3.11 (sept, J = 6.8 Hz, 4H), 1.35 (d, J = 12.0 Hz, 12H), 1.32 (d, J = 6.9 Hz, 12H). 13C NMR (75 MHz, DMSO-d6) δ 208.19, 146.51, 134.70, 134.45, 132.61, 130.54, 129.82, 129.78, 129.57, 127.79, 125.98, 124.34, 102.07, 53.76, 28.21, 25.04, 23.65.
[Au(SIPr)(C≡C(2-pyridyl] (4). The synthesis of this compound required the previous deprotonation of the alkyne. The ethynylpyridine (27 mg, 0.262 mmol) was dissolved in 5 mL of a solution of 0.11 M of sodium methoxide in MeOH and maintained for 30 min with stirring. [Au(SIPr)Cl] (80 mg, 0.129 mmol) was added to the solution, and the mixture was reacted for 24 h. Following work-up was analogous to that described in the general procedure. Yield 79 %. Anal. Calcd. For C34H42N3Au: C, 59.21; H, 6.15; N, 6.09. Found: C, 58.92; H, 6.16; N, 5.86. Ir (KBr, cm−1): 2960(νC–H), 2118(νC≡C), 1583(νCNpyridine), 1490(νCCarom), 1460(ν−Carom), 1275(νC–N). 1H NMR (300 MHz, DMSO-d6) δ 8.29–8.25 (m, 1H), 7.55–7.48 (m, 1H), 7.48–7.41 (m, 2H), 7.34 (d, J = 7.6 Hz, 4H), 7.09–7.03 (m, 2H), 4.09 (s, 4H), 3.11 (sept, J = 6.8 Hz, 4H), 1.36 (d, J = 6.8 Hz, 12H), 1.32 (d, J = 6.8 Hz, 12H). 13C NMR (75 MHz, DMSO-d6) δ 208.20, 149.15, 146.50, 144.58, 135.77, 134.38, 133.21, 129.59, 126.23, 124.31, 120.94, 103.69, 53.74, 28.21, 25.05, 23.66.
[Au(SIPr)(C≡C(4-bromophenyl] (5). Yield 29 %. Anal. Calcd. For C35H42N2AuBr: C, 54.76; H, 5.53; N, 3.65. Found: C, 54.81; H, 5.57; N, 3.53. Ir (KBr, cm−1): 2961(νC–H), 2112(νC≡C), 1489(νCCarom), 1458(νCCarom), 1272(νC–N), 619(νC–Br). 1H NMR (300 MHz, DMSO-d6) δ 7.45 (dd, J = 8.5, 6.9 Hz, 2H), 7.33 (d, J = 7.7 Hz, 4H), 7.31–7.26 (m, 2H), 7.03–6.96 (m, 2H), 4.07 (s, 4H), 3.10 (sept, J = 6.9 Hz, 4H), 1.35 (d, J = 6.9 Hz, 12H), 1.31 (d, J = 6.9 Hz, 12H). 13C NMR (75 MHz, DMSO-d6) δ 208.28, 146.50, 134.45, 134.27, 133.10, 130.91, 129.55, 125.08, 124.33, 118.84, 102.41, 53.74, 28.20, 25.03, 23.65.

4. Conclusions

The reaction of [Au(SIPr)Cl] with different arylalkynes in the presence of a base yielded complexes 15, which were characterized and whose cyclic voltammetry was explored both under argon and under CO2 atmosphere. In the cyclic voltammetry measurements under argon atmosphere, complexes 24 display irreversible reduction at values ranging between −2.79 and −2.91 V referenced to the ferrocenium/ferrocene couple. Complex 1 does not show this reduction, and this fact can be understood by comparing the energies of the LUMOs of the complexes. The energy of the LUMO of complex 1 is higher than the energy of the LUMOs of the other compounds, and consequently complex 1 is expected to experience a reduction at a lower potential than that which should appear outside of the measurement window of the experiment. The DFT calculations show that the LUMO is mainly contributed to by the atomic orbitals of atoms of the aryl–C≡C fragments with little participation from the atomic gold orbitals, and the reduction takes place mainly in the aryl–C≡C fragments. The cyclic voltammetry measurements performed under carbon dioxide atmosphere showed an increase in the faradaic current of the reduction wave of these complexes, indicating that there is activity towards the CO2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29174081/s1, Figures S1–S3: intermolecular interactions in the crystal structure; Figures S4–S8: Cyclic voltammograms of complexes 15; Figures S9–S12: energy levels and frontier molecular orbitals of complexes 1, 35; Table S1: Composition of the molecular orbitals for complexes 15; Figures S13–S16; 1H and 13C NMR spectra for complexes 25.

Author Contributions

A.R.-R. and Á.Y.: synthetic procedures, characterization of products, and cyclic voltammetry experiments; G.G.-H.: electrochemical interpretation and partial writing; T.T.: reviewing and editing, funding acquisition; J.V.C.-V.: DFT calculations, writing—reviewing and editing and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Ministerio de Ciencia, Innovación y Universidades of Spain (Grant PID2022-142318NB-I00) and Junta de Castilla y León, the Spanish Ministerio de Ciencia e Innovación MICIN, and the European Union Next Generation EU/PRTR (MR6NP3).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

We also thank Marta Masilla from the Parque Científico y Tecnológico de la Universidad de Burgos for X-ray acquisition.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mulks, F.F. Gold Carbene Complexes and beyond: New Avenues in Gold(I)—Carbon Coordination Chemistry. Gold Bull. 2022, 55, 1–13. [Google Scholar] [CrossRef]
  2. Lin, I.J.; Vasam, C.S. Review of Gold(I) N−Heterocyclic Carbenes. Can. J. Chem. 2005, 83, 812–825. [Google Scholar] [CrossRef]
  3. Mora, M.; Gimeno, M.C.; Visbal, R. Recent Advances in Gold–NHC Complexes with Biological Properties. Chem. Soc. Rev. 2019, 48, 447–462. [Google Scholar] [CrossRef] [PubMed]
  4. Hussaini, S.Y.; Haque, R.A.; Razali, M.R. Recent Progress in Silver(I)-, Gold(I)/(III)- and Palladium(II)-N-Heterocyclic Carbene Complexes: A Review towards Biological Perspectives. J. Organomet. Chem. 2019, 882, 96–111. [Google Scholar] [CrossRef]
  5. Mertens, R.T.; Gukathasan, S.; Arojojoye, A.S.; Olelewe, C.; Awuah, S.G. Next Generation Gold Drugs and Probes: Chemistry and Biomedical Applications. Chem. Rev. 2023, 123, 6612–6667. [Google Scholar] [CrossRef]
  6. Zhang, Y.-F.; Yin, Y.-K.; Zhang, H.; Han, Y.-F. Metal N-Heterocyclic Carbene Complexes as Potential Metallodrugs in Antitumor Therapy. Coord. Chem. Rev. 2024, 514, 215941. [Google Scholar] [CrossRef]
  7. Tialiou, A.; Chin, J.; Keppler, B.K.; Reithofer, M.R. Current Developments of N-Heterocyclic Carbene Au(I)/Au(III) Complexes toward Cancer Treatment. Biomedicines 2022, 10, 1417. [Google Scholar] [CrossRef]
  8. Lu, Y.; Ma, X.; Chang, X.; Liang, Z.; Lv, L.; Shan, M.; Lu, Q.; Wen, Z.; Gust, R.; Liu, W. Recent Development of Gold(I) and Gold(III) Complexes as Therapeutic Agents for Cancer Diseases. Chem. Soc. Rev. 2022, 51, 5518–5556. [Google Scholar] [CrossRef]
  9. Yeo, C.I.; Goh, C.H.P.; Tiekink, E.R.T.; Chew, J. Antibiotics: A “GOLDen” Promise? Coord. Chem. Rev. 2024, 500, 215429. [Google Scholar] [CrossRef]
  10. Scattolin, T.; Tonon, G.; Botter, E.; Guillet, S.G.; Tzouras, N.V.; Nolan, S.P. Gold(I)- N -Heterocyclic Carbene Synthons in Organometallic Synthesis. Chem. Eur. J. 2023, 29, e202301961. [Google Scholar] [CrossRef]
  11. Marion, N.; Nolan, S.P. N-Heterocyclic Carbenes in Gold Catalysis. Chem. Soc. Rev. 2008, 37, 1776. [Google Scholar] [CrossRef] [PubMed]
  12. Gade, A.B.; Urvashi; Patil, N.T. Asymmetric Gold Catalysis Enabled by Specially Designed Ligands. Org. Chem. Front. 2024, 11, 1858–1895. [Google Scholar] [CrossRef]
  13. Michalak, M.; Kośnik, W. Chiral N-Heterocyclic Carbene Gold Complexes: Synthesis and Applications in Catalysis. Catalysts 2019, 9, 890. [Google Scholar] [CrossRef]
  14. Jiang, J.; Wong, M. Recent Advances in the Development of Chiral Gold Complexes for Catalytic Asymmetric Catalysis. Chem.An Asian J. 2021, 16, 364–377. [Google Scholar] [CrossRef]
  15. Mato, M.; Franchino, A.; García-Morales, C.; Echavarren, A.M. Gold-Catalyzed Synthesis of Small Rings. Chem. Rev. 2021, 121, 8613–8684. [Google Scholar] [CrossRef]
  16. Praveen, C.; Dupeux, A.; Michelet, V. Catalytic Gold Chemistry: From Simple Salts to Complexes for Regioselective C–H Bond Functionalization. Chem. A Eur. J. 2021, 27, 10495–10532. [Google Scholar] [CrossRef]
  17. Zhao, Q.; Meng, G.; Nolan, S.P.; Szostak, M. N-Heterocyclic Carbene Complexes in C–H Activation Reactions. Chem. Rev. 2020, 120, 1981–2048. [Google Scholar] [CrossRef]
  18. Nijamudheen, A.; Datta, A. Gold-Catalyzed Cross-Coupling Reactions: An Overview of Design Strategies, Mechanistic Studies, and Applications. Chem. Eur. J. 2020, 26, 1442–1487. [Google Scholar] [CrossRef]
  19. Gao, P.; Szostak, M. Hydration Reactions Catalyzed by Transition Metal–NHC (NHC = N-Heterocyclic Carbene) Complexes. Coord. Chem. Rev. 2023, 485, 215110. [Google Scholar] [CrossRef]
  20. Crochet, P.; Cadierno, V. Gold Complexes with Hydrophilic N-Heterocyclic Carbene Ligands and Their Contribution to Aqueous-Phase Catalysis. Catalysts 2023, 13, 436. [Google Scholar] [CrossRef]
  21. Dang, H.; Guan, B.; Chen, J.; Ma, Z.; Chen, Y.; Zhang, J.; Guo, Z.; Chen, L.; Hu, J.; Yi, C.; et al. Research Status, Challenges, and Future Prospects of Carbon Dioxide Reduction Technology. Energy Fuels 2024, 38, 4836–4880. [Google Scholar] [CrossRef]
  22. Almomani, F.; Abdelbar, A.; Ghanimeh, S. A Review of the Recent Advancement of Bioconversion of Carbon Dioxide to Added Value Products: A State of the Art. Sustainability 2023, 15, 10438. [Google Scholar] [CrossRef]
  23. Akash, S.; Sivaprakash, B.; Rajamohan, N.; Vo, D.-V.N. Biotechnology to Convert Carbon Dioxide into Biogas, Bioethanol, Bioplastic and Succinic Acid Using Algae, Bacteria and Yeast: A Review. Environ. Chem. Lett. 2023, 21, 1477–1497. [Google Scholar] [CrossRef]
  24. Kondaveeti, S.; Abu-Reesh, I.M.; Mohanakrishna, G.; Bulut, M.; Pant, D. Advanced Routes of Biological and Bio-Electrocatalytic Carbon Dioxide (CO2) Mitigation Toward Carbon Neutrality. Front. Energy Res. 2020, 8, 94. [Google Scholar] [CrossRef]
  25. Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.; Hille, R.; Kenis, P.J.A.; et al. Frontiers, Opportunities, and Challenges in Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 6621–6658. [Google Scholar] [CrossRef]
  26. Hakami, O. Thermocatalytic and Solar Thermochemical Carbon Dioxide Utilization to Solar Fuels and Chemicals: A Review. Int. J. Energy Res. 2022, 46, 19929–19960. [Google Scholar] [CrossRef]
  27. Abanades, S. Redox Cycles, Active Materials, and Reactors Applied to Water and Carbon Dioxide Splitting for Solar Thermochemical Fuel Production: A Review. Energies 2022, 15, 7061. [Google Scholar] [CrossRef]
  28. Wei, L.; Pan, Z.; Shi, X.; Esan, O.C.; Li, G.; Qi, H.; Wu, Q.; An, L. Solar-Driven Thermochemical Conversion of H2O and CO2 into Sustainable Fuels. iScience 2023, 26, 108127. [Google Scholar] [CrossRef]
  29. Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A. Fundamentals and Applications of Photocatalytic CO2 Methanation. Nat. Commun. 2019, 10, 3169. [Google Scholar] [CrossRef]
  30. Canivet, J.; Wisser, F.M. Metal–Organic Framework Catalysts for Solar Fuels: Light-Driven Conversion of Carbon Dioxide into Formic Acid. ACS Appl. Energy Mater. 2023, 6, 9027–9043. [Google Scholar] [CrossRef]
  31. Pachaiappan, R.; Rajendran, S.; Senthil Kumar, P.; Vo, D.-V.N.; Hoang, T.K. A Review of Recent Progress on Photocatalytic Carbon Dioxide Reduction into Sustainable Energy Products Using Carbon Nitride. Chem. Eng. Res. Des. 2022, 177, 304–320. [Google Scholar] [CrossRef]
  32. Van Le, Q.; Nguyen, V.-H.; Nguyen, T.D.; Sharma, A.; Rahman, G.; Nguyen, D.L.T. Light-Driven Reduction of Carbon Dioxide: Altering the Reaction Pathways and Designing Photocatalysts toward Value-Added and Renewable Fuels. Chem. Eng. Sci. 2021, 237, 116547. [Google Scholar] [CrossRef]
  33. Han, J.; Bai, X.; Xu, X.; Bai, X.; Husile, A.; Zhang, S.; Qi, L.; Guan, J. Advances and Challenges in the Electrochemical Reduction of Carbon Dioxide. Chem. Sci. 2024, 15, 7870–7907. [Google Scholar] [CrossRef]
  34. Zhang, X.; Guo, S.-X.; Gandionco, K.A.; Bond, A.M.; Zhang, J. Electrocatalytic Carbon Dioxide Reduction: From Fundamental Principles to Catalyst Design. Mater. Today Adv. 2020, 7, 100074. [Google Scholar] [CrossRef]
  35. Liu, B.; Wei, Z.; Ding, J.; Wang, Z.; Wang, A.; Zhang, L.; Liu, Y.; Guo, Y.; Yang, X.; Zhai, Y. Enhanced Electrochemical CO2 Reduction to Formate over Phosphate-Modified In: Water Activation and Active Site Tuning. Angew. Chemie Int. Ed. 2024, 136, e202402070. [Google Scholar] [CrossRef]
  36. Zhang, F.; Zhang, H.; Liu, Z. Recent Advances in Electrochemical Reduction of CO2. Curr. Opin. Green Sustain. Chem. 2019, 16, 77–84. [Google Scholar] [CrossRef]
  37. Kim, D.; Sakimoto, K.K.; Hong, D.; Yang, P. Artificial Photosynthesis for Sustainable Fuel and Chemical Production. Angew. Chem. Int. Ed. 2015, 54, 3259–3266. [Google Scholar] [CrossRef]
  38. Al-Tamreh, S.A.; Ibrahim, M.H.; El-Naas, M.H.; Vaes, J.; Pant, D.; Benamor, A.; Amhamed, A. Electroreduction of Carbon Dioxide into Formate: A Comprehensive Review. ChemElectroChem 2021, 8, 3207–3220. [Google Scholar] [CrossRef]
  39. Lee, M.-Y.; Park, K.T.; Lee, W.; Lim, H.; Kwon, Y.; Kang, S. Current Achievements and the Future Direction of Electrochemical CO2 Reduction: A Short Review. Crit. Rev. Environ. Sci. Technol. 2020, 50, 769–815. [Google Scholar] [CrossRef]
  40. Zhang, J.; Cai, W.; Hu, F.X.; Yang, H.; Liu, B. Recent Advances in Single Atom Catalysts for the Electrochemical Carbon Dioxide Reduction Reaction. Chem. Sci. 2021, 12, 6800–6819. [Google Scholar] [CrossRef]
  41. Duarah, P.; Haldar, D.; Yadav, V.; Purkait, M.K. Progress in the Electrochemical Reduction of CO2 to Formic Acid: A Review on Current Trends and Future Prospects. J. Environ. Chem. Eng. 2021, 9, 106394. [Google Scholar] [CrossRef]
  42. Jin, S.; Hao, Z.; Zhang, K.; Yan, Z.; Chen, J. Advances and Challenges for the Electrochemical Reduction of CO2 to CO: From Fundamentals to Industrialization. Angew. Chem. Int. Ed. 2021, 60, 20627–20648. [Google Scholar] [CrossRef]
  43. Zhang, S.; Fan, Q.; Xia, R.; Meyer, T.J. CO2 Reduction: From Homogeneous to Heterogeneous Electrocatalysis. Acc. Chem. Res. 2020, 53, 255–264. [Google Scholar] [CrossRef]
  44. Feng, D.-M.; Zhu, Y.-P.; Chen, P.; Ma, T.-Y. Recent Advances in Transition-Metal-Mediated Electrocatalytic CO2 Reduction: From Homogeneous to Heterogeneous Systems. Catalysts 2017, 7, 373. [Google Scholar] [CrossRef]
  45. Kinzel, N.W.; Werlé, 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]
  46. Boogaerts, I.I.F.; Nolan, S.P. Carboxylation of C−H Bonds Using N -Heterocyclic Carbene Gold(I) Complexes. J. Am. Chem. Soc. 2010, 132, 8858–8859. [Google Scholar] [CrossRef]
  47. D’Amato, A.; Sirignano, M.; Russo, S.; Troiano, R.; Mariconda, A.; Longo, P. Recent Advances in N-Heterocyclic Carbene Coinage Metal Complexes in A3-Coupling and Carboxylation Reaction. Catalysts 2023, 13, 811. [Google Scholar] [CrossRef]
  48. Hase, S.; Kayaki, Y.; Ikariya, T. NHC–Gold(I) Complexes as Effective Catalysts for the Carboxylative Cyclization of Propargylamines with Carbon Dioxide. Organometallics 2013, 32, 5285–5288. [Google Scholar] [CrossRef]
  49. Wen, D.; Zheng, Q.; Yang, S.; Zhu, H.; Tu, T. Direct Knitting Boosts the Stability and Catalytic Activity of NHC-Au Complexes towards Valorization of SO2 and CO2. J. Catal. 2023, 418, 64–69. [Google Scholar] [CrossRef]
  50. Collado, A.; Gómez-Suárez, A.; Webb, P.B.; Kruger, H.; Bühl, M.; Cordes, D.B.; Slawin, A.M.Z.; Nolan, S.P. Trapping Atmospheric CO2 with Gold. Chem. Commun. 2014, 50, 11321–11324. [Google Scholar] [CrossRef]
  51. Zhang, C.; Ding, M.; Ren, Y.; Ma, A.; Yin, Z.; Ma, X.; Wang, S. The Smallest Superatom Au4(PPh3)4I2 with Two Free Electrons: Synthesis, Structure Analysis, and Electrocatalytic Conversion of CO2 to CO. Nanoscale Adv. 2023, 5, 3287–3292. [Google Scholar] [CrossRef] [PubMed]
  52. Levchenko, T.I.; Yi, H.; Aloisio, M.D.; Dang, N.K.; Gao, G.; Sharma, S.; Dinh, C.-T.; Crudden, C.M. Electrocatalytic CO2 Reduction with Atomically Precise Au13 Nanoclusters: Effect of Ligand Shell on Catalytic Performance. ACS Catal. 2024, 14, 4155–4163. [Google Scholar] [CrossRef]
  53. Sun, Y.; Liu, X.; Xiao, K.; Zhu, Y.; Chen, M. Active-Site Tailoring of Gold Cluster Catalysts for Electrochemical CO2 Reduction. ACS Catal. 2021, 11, 11551–11560. [Google Scholar] [CrossRef]
  54. Gao, Z.-H.; Wei, K.; Wu, T.; Dong, J.; Jiang, D.; Sun, S.; Wang, L.-S. A Heteroleptic Gold Hydride Nanocluster for Efficient and Selective Electrocatalytic Reduction of CO2 to CO. J. Am. Chem. Soc. 2022, 144, 5258–5262. [Google Scholar] [CrossRef] [PubMed]
  55. Sun, K.; Shi, Y.; Li, H.; Shan, J.; Sun, C.; Wu, Z.; Ji, Y.; Wang, Z. Efficient CO2 Electroreduction via Au-Complex Derived Carbon Nanotube Supported Au Nanoclusters. ChemSusChem 2021, 14, 4929–4935. [Google Scholar] [CrossRef]
  56. Seong, H.; Efremov, V.; Park, G.; Kim, H.; Yoo, J.S.; Lee, D. Atomically Precise Gold Nanoclusters as Model Catalysts for Identifying Active Sites for Electroreduction of CO2. Angew. Chem. Int. Ed. 2021, 60, 14563–14570. [Google Scholar] [CrossRef]
  57. Tang, L.; Luo, Y.; Ma, X.; Wang, B.; Ding, M.; Wang, R.; Wang, P.; Pei, Y.; Wang, S. Poly-Hydride [AuI7(PPh3)7H5](SbF6)2 Cluster Complex: Structure, Transformation, and Electrocatalytic CO2 Reduction Properties. Angew. Chem. Int. Ed. 2023, 62, e202402070. [Google Scholar] [CrossRef]
  58. Albright, E.L.; Malola, S.; Jacob, S.I.; Yi, H.; Takano, S.; Mimura, K.; Tsukuda, T.; Häkkinen, H.; Nambo, M.; Crudden, C.M. Enantiopure Chiral Au13 Nanoclusters Stabilized by Ditopic N -Heterocyclic Carbenes: Synthesis, Characterization, and Electrocatalytic Reduction of CO2. Chem. Mater. 2024, 36, 1279–1289. [Google Scholar] [CrossRef]
  59. Scattolin, T.; Tzouras, N.V.; Falivene, L.; Cavallo, L.; Nolan, S.P. Using Sodium Acetate for the Synthesis of [Au(NHC)X] Complexes. Dalt. Trans. 2020, 49, 9694–9700. [Google Scholar] [CrossRef]
  60. Danovich, D.; Shaik, S.; Neese, F.; Echeverría, J.; Aullón, G.; Alvarez, S. Understanding the Nature of the CH···HC Interactions in Alkanes. J. Chem. Theory Comput. 2013, 9, 1977–1991. [Google Scholar] [CrossRef]
  61. Hamze, R.; Shi, S.; Kapper, S.C.; Muthiah Ravinson, D.S.; Estergreen, L.; Jung, M.-C.; Tadle, A.C.; Haiges, R.; Djurovich, P.I.; Peltier, J.L.; et al. “Quick-Silver” from a Systematic Study of Highly Luminescent, Two-Coordinate, d10 Coinage Metal Complexes. J. Am. Chem. Soc. 2019, 141, 8616–8626. [Google Scholar] [CrossRef]
  62. Hamze, R.; Idris, M.; Muthiah Ravinson, D.S.; Jung, M.C.; Haiges, R.; Djurovich, P.I.; Thompson, M.E. Highly Efficient Deep Blue Luminescence of 2-Coordinate Coinage Metal Complexes Bearing Bulky NHC Benzimidazolyl Carbene. Front. Chem. 2020, 8, 401. [Google Scholar] [CrossRef]
  63. Ruiz-Zambrana, C.; Poyatos, M.; Peris, E. A Redox-Switchable Gold(I) Complex for the Hydroamination of Acetylenes: A Convenient Way for Studying Ligand-Derived Electronic Effects. ACS Catal. 2022, 12, 4465–4472. [Google Scholar] [CrossRef]
  64. Huynh, H.V.; Guo, S.; Wu, W. Detailed Structural, Spectroscopic, and Electrochemical Trends of Halido Mono- and Bis(NHC) Complexes of Au(I) and Au(III). Organometallics 2013, 32, 4591–4600. [Google Scholar] [CrossRef]
  65. McCall, R.; Miles, M.; Lascuna, P.; Burney, B.; Patel, Z.; Sidoran, K.J.; Sittaramane, V.; Kocerha, J.; Grossie, D.A.; Sessler, J.L.; et al. Dual Targeting of the Cancer Antioxidant Network with 1,4-Naphthoquinone Fused Gold(I) N-Heterocyclic Carbene Complexes. Chem. Sci. 2017, 8, 5918–5929. [Google Scholar] [CrossRef]
  66. Klenk, S.; Rupf, S.; Suntrup, L.; van der Meer, M.; Sarkar, B. The Power of Ferrocene, Mesoionic Carbenes, and Gold: Redox-Switchable Catalysis. Organometallics 2017, 36, 2026–2035. [Google Scholar] [CrossRef]
  67. Pažický, M.; Loos, A.; Ferreira, M.J.; Serra, D.; Vinokurov, N.; Rominger, F.; Jäkel, C.; Hashmi, A.S.K.; Limbach, M. Synthesis, Reactivity, and Electrochemical Studies of Gold(I) and Gold(III) Complexes Supported by N-Heterocyclic Carbenes and Their Application in Catalysis. Organometallics 2010, 29, 4448–4458. [Google Scholar] [CrossRef]
  68. Méndez-Hernández, D.D.; Tarakeshwar, P.; Gust, D.; Moore, T.A.; Moore, A.L.; Mujica, V. Simple and Accurate Correlation of Experimental Redox Potentials and DFT-Calculated HOMO/LUMO Energies of Polycyclic Aromatic Hydrocarbons. J. Mol. Model. 2013, 19, 2845–2848. [Google Scholar] [CrossRef]
  69. Conradie, J. A Frontier Orbital Energy Approach to Redox Potentials. J. Phys. Conf. Ser. 2015, 633, 012045. [Google Scholar] [CrossRef]
  70. Athanasopoulos, E.; Conradie, J. Substituent Effect on the Oxidation and Reduction of Electronically Altered Iron(II)-Terpyridine Derivatives–DFT Study. Inorg. Chim. Acta 2023, 555, 121599. [Google Scholar] [CrossRef]
  71. Chiyindiko, E.; Langner, E.H.G.; Conradie, J. Electrochemical Behaviour of Copper(II) Complexes Containing 2-Hydroxyphenones. Electrochim. Acta 2022, 424, 140629. [Google Scholar] [CrossRef]
  72. Liyanage, N.P.; Dulaney, H.A.; Huckaba, A.J.; Jurss, J.W.; Delcamp, J.H. Electrocatalytic Reduction of CO2 to CO With Re-Pyridyl-NHCs: Proton Source Influence on Rates and Product Selectivities. Inorg. Chem. 2016, 55, 6085–6094. [Google Scholar] [CrossRef]
  73. Stanton, C.J.; Machan, C.W.; Vandezande, J.E.; Jin, T.; Majetich, G.F.; Schaefer, H.F.; Kubiak, C.P.; Li, G.; Agarwal, J. Re(I) NHC Complexes for Electrocatalytic Conversion of CO2. Inorg. Chem. 2016, 55, 3136–3144. [Google Scholar] [CrossRef]
  74. 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] [PubMed]
  75. Yang, Y.; Zhang, Z.; Chang, X.; Zhang, Y.-Q.; Liao, R.-Z.; Duan, L. Highly Active Manganese-Based CO2 Reduction Catalysts with Bulky NHC Ligands: A Mechanistic Study. Inorg. Chem. 2020, 59, 10234–10242. [Google Scholar] [CrossRef] [PubMed]
  76. Fernández, S.; Franco, F.; Martínez Belmonte, M.; Friães, S.; Royo, B.; Luis, J.M.; Lloret-Fillol, J. Decoding the CO2 Reduction Mechanism of a Highly Active Organometallic Manganese Electrocatalyst: Direct Observation of a Hydride Intermediate and Its Implications. ACS Catal. 2023, 13, 10375–10385. [Google Scholar] [CrossRef]
  77. Gonell, S.; Lloret-Fillol, J.; Miller, A.J.M. An Iron Pyridyl-Carbene Electrocatalyst for Low Overpotential CO2 Reduction to CO. ACS Catal. 2021, 11, 615–626. [Google Scholar] [CrossRef]
  78. Gonell, S.; Assaf, E.A.; Lloret-Fillol, J.; Miller, A.J.M. An Iron Bis(Carbene) Catalyst for Low Overpotential CO2 Electroreduction to CO: Mechanistic Insights from Kinetic Zone Diagrams, Spectroscopy, and Theory. ACS Catal. 2021, 11, 15212–15222. [Google Scholar] [CrossRef]
  79. Gonell, S.; Assaf, E.A.; Duffee, K.D.; Schauer, C.K.; Miller, A.J.M. Kinetics of the Trans Effect in Ruthenium Complexes Provide Insight into the Factors That Control Activity and Stability in CO2 Electroreduction. J. Am. Chem. Soc. 2020, 142, 8980–8999. [Google Scholar] [CrossRef]
  80. Liu, Y.; Fan, X.; Nayak, A.; Wang, Y.; Shan, B.; Quan, X.; Meyer, T.J. Steering CO2 Electroreduction toward Ethanol Production by a Surface-Bound Ru Polypyridyl Carbene Catalyst on N-Doped Porous Carbon. Proc. Natl. Acad. Sci. USA 2019, 116, 26353–26358. [Google Scholar] [CrossRef]
  81. Shirley, H.; Figgins, M.T.; Boudreaux, C.M.; Liyanage, N.P.; Lamb, R.W.; Webster, C.E.; Papish, E.T.; Delcamp, J.H. Impact of the Dissolved Anion on the Electrocatalytic Reduction of CO2 to CO with Ruthenium CNC Pincer Complexes. ChemCatChem 2020, 12, 4879–4885. [Google Scholar] [CrossRef]
  82. Mudge, M.N.; Bhadbhade, M.; Ball, G.E.; Colbran, S.B. Ruthenium(II) Complexes of a Xanthene-Spanned Dicarbene Ligand. Inorg. Chem. 2023, 62, 18901–18914. [Google Scholar] [CrossRef]
  83. Assaf, E.A.; Gonell, S.; Chen, C.-H.; Miller, A.J.M. Accessing and Photo-Accelerating Low-Overpotential Pathways for CO2 Reduction: A Bis-Carbene Ruthenium Terpyridine Catalyst. ACS Catal. 2022, 12, 12596–12606. [Google Scholar] [CrossRef]
  84. Su, X.; McCardle, K.M.; Chen, L.; Panetier, J.A.; Jurss, J.W. Robust and Selective Cobalt Catalysts Bearing Redox-Active Bipyridyl-N-Heterocyclic Carbene Frameworks for Electrochemical CO2 Reduction in Aqueous Solutions. ACS Catal. 2019, 9, 7398–7408. [Google Scholar] [CrossRef]
  85. Su, X.; McCardle, K.M.; Panetier, J.A.; Jurss, J.W. Electrocatalytic CO2 Reduction with Nickel Complexes Supported by Tunable Bipyridyl-N-Heterocyclic Carbene Donors: Understanding Redox-Active Macrocycles. Chem. Commun. 2018, 54, 3351–3354. [Google Scholar] [CrossRef]
  86. DeLuca, E.E.; Xu, Z.; Lam, J.; Wolf, M.O. Improved Electrocatalytic CO2 Reduction with Palladium Bis(NHC) Pincer Complexes Bearing Cationic Side Chains. Organometallics 2019, 38, 1330–1343. [Google Scholar] [CrossRef]
  87. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  88. Lee, C.T.; Yang, W.T.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron-Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  89. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2019. [Google Scholar]
  90. Hay, P.J.; Wadt, W.R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299–310. [Google Scholar] [CrossRef]
  91. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self—Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  92. Francl, M.M.; Pietro, W.J.; Hehre, W.J.; Binkley, J.S.; Gordon, M.S.; DeFrees, D.J.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XXIII. A Polarization-Type Basis Set for Second-Row Elements. J. Chem. Phys. 1982, 77, 3654–3665. [Google Scholar] [CrossRef]
  93. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  94. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  95. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef] [PubMed]
  96. SAINT+. Area-Detector Integration Program; Bruker AXS Inc.: Madison, WI, USA, 2004. [Google Scholar]
  97. SADABS-2016/2; Bruker AXS Inc.: Madison, WI, USA, 2016.
  98. SHELXTL-NT; Structure Determination Package. Bruker AXS Inc.: Madison, WI, USA, 2001.
Scheme 1. General procedure of synthesis.
Scheme 1. General procedure of synthesis.
Molecules 29 04081 sch001
Figure 1. Molecular structure and atomic labeling of complex 2. Hydrogen atoms are omitted to see the structures more clearly. Thermal ellipsoids are 30%. The code of colors of the atoms is: Carbon (grey), Nitrogen (blue), Gold (yellow).
Figure 1. Molecular structure and atomic labeling of complex 2. Hydrogen atoms are omitted to see the structures more clearly. Thermal ellipsoids are 30%. The code of colors of the atoms is: Carbon (grey), Nitrogen (blue), Gold (yellow).
Molecules 29 04081 g001
Figure 2. Fragment of cyclic voltammogram of complex 2 (5·10−4 M, glassy carbon working electrode dish with 3.0 mm diameter, dry MeCN, 0.1 M Bu4NPF6, scan rate 100 mV·s−1), recorded under argon atmosphere (blue line) and CO2 atmosphere (orange line). The arrow indicates the scan direction. The complete cyclic voltammogram can be seen in Figure S5 of the Supplementary Materials.
Figure 2. Fragment of cyclic voltammogram of complex 2 (5·10−4 M, glassy carbon working electrode dish with 3.0 mm diameter, dry MeCN, 0.1 M Bu4NPF6, scan rate 100 mV·s−1), recorded under argon atmosphere (blue line) and CO2 atmosphere (orange line). The arrow indicates the scan direction. The complete cyclic voltammogram can be seen in Figure S5 of the Supplementary Materials.
Molecules 29 04081 g002
Figure 3. Calculated energy levels for the FMOs of complexes 15.
Figure 3. Calculated energy levels for the FMOs of complexes 15.
Molecules 29 04081 g003
Figure 4. Energy levels and isosurface contour plots (0.03 au) for complex 2.
Figure 4. Energy levels and isosurface contour plots (0.03 au) for complex 2.
Molecules 29 04081 g004
Table 1. Crystallographic data for 2.
Table 1. Crystallographic data for 2.
Empirical FormulaC43H47AuN2Formula Weight788.79
temperature/K240crystal systemorthorhombic
space groupP212121
a13.1684 (4)α/deg90
b16.4356 (5)β/deg90
c38.3544 (12)γ/deg90
volume/Å38301.1 (4)Z8
ρcalc g/cm31.262μ/mm−16.856
F (000)3184.0radiationCu Kα (λ = 1.54178)
2Θ range for data collection/deg4.07 to 72.23index ranges−15 ≤ h ≤ 16−20 ≤ k ≤ 20−47 ≤ l ≤ 45
reflections collected16064goodness-of-fit on F21.035
final R indexes [I ≥ 2σ (I)]R1 = 0.0288final [all data]R1 = 0.0328
wR2 = 0.0700 wR2 = 0.0730
largest diff. peak/hole/e Å−30.732/−0.584CCDC number2,363,526
Table 2. Selected bond lengths (Å), angles (deg) and torsion angles (deg) for 2.
Table 2. Selected bond lengths (Å), angles (deg) and torsion angles (deg) for 2.
XRDDFT XRDDFT
Au2–C602.036 (6)2.045Au2–C592.003 (7)2.008
Au1–C172.033 (6)2.045Au1–C161.983 (6)2.008
N3–C601.311 (8)1.336N3–C611.471 (8)1.479
N2–C171.329 (8)1.337N2–C191.469 (8)1.479
N4–C601.331 (8)1.336N4–C621.463 (8)1.479
N1–C171.320 (8)1.337N1–C181.470 (8)1.479
C61–C621.527 (10)1.542C14–C151.433 (9)1.426
C16–C151.207 (9)1.228C19–C181.502 (10)1.539
C57–C581.442 (9)1.426C58–C591.192 (9)1.228
C59–Au2–C60176.1 (3)179.1C16–Au1–C17173.8 (3)179.0
C60–N3–C61113.4 (5)113.1C17–N2–C19111.7 (5)112.8
N3–C60–Au2126.7 (4)125.0N3–C60–N4108.7 (5)109.1
N4–C60–Au2124.5 (4)126.0C60–N4–C62112.7 (5)113.0
N2–C17–Au1126.5 (5)125.7N1–C17–Au1123.5 (5)125.3
N1–C17–N2109.5 (5)109.0C17–N1–C18112.7 (5)112.8
C58–C59–Au2171.1 (6)178.6N3–C61–C62101.6 (5)102.3
C15–C16–Au1168.2 (6)178.3N2–C19–C18103.5 (5)102.2
N4–C62–C61102.2 (5)102.4N1–C18–C19102.5 (5)102.2
C60–N3–C75–C7691.7 (8)87.7C60–N3–C75–C7988.5 (8)91.6
C60–N4–C63–C6884.2 (8)87.2C60–N4–C63–C6497.1 (8)92.1
C17–N2–C20–C2596.6 (8)95.7C17–N2–C20–C2182.3 (9)83.9
C17–N1–C32–C3382.2 (9)96.9C17–N1–C32–C3798.8 (8)82.6
Table 3. Cyclic voltammetry data for compounds 15 in MeCN and Bu4PF6 supporting electrolyte, using the redox system ferrocenium/ferrocene as internal reference a.
Table 3. Cyclic voltammetry data for compounds 15 in MeCN and Bu4PF6 supporting electrolyte, using the redox system ferrocenium/ferrocene as internal reference a.
Compound E p red (V) Ar biAr(μA) c E p red (V) CO2diCO2(μA) e Ratio   f   i C O 2 i A r
1 −2.99−92.74
2−2.87(ir)−38.36−2.88−95.582.49
3−2.90(ir)−29.11−3.04−56.501.94
4−2.91(ir)−46.51−1.40/−3.00−35.5/−82.491.77
5−2.79(ir)−26.46/−1.37/−3.04−18.49/−68.942.61
a The reduction potential mean value observed for ferrocenium/ferrocene (Fc+/Fc), used as an internal calibrant under the employed experimental conditions, was E° = 0.476 V vs. the AgCl/Ag (3 M KCl) electrode. b Cathodic peaks observed under argon. c Maximum registered cathodic current under argon. d Cathodic peaks observed under CO2. e Maximum registered cathodic current under CO2. f Ratio between the faradaic currents observed under CO2 (iCO2) and under argon (iAr).
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

Rodríguez-Rubio, A.; Yuste, Á.; Torroba, T.; García-Herbosa, G.; Cuevas-Vicario, J.V. Synthesis and Electrochemical Study of Gold(I) Carbene Complexes. Molecules 2024, 29, 4081. https://doi.org/10.3390/molecules29174081

AMA Style

Rodríguez-Rubio A, Yuste Á, Torroba T, García-Herbosa G, Cuevas-Vicario JV. Synthesis and Electrochemical Study of Gold(I) Carbene Complexes. Molecules. 2024; 29(17):4081. https://doi.org/10.3390/molecules29174081

Chicago/Turabian Style

Rodríguez-Rubio, Andrea, Álvaro Yuste, Tomás Torroba, Gabriel García-Herbosa, and José V. Cuevas-Vicario. 2024. "Synthesis and Electrochemical Study of Gold(I) Carbene Complexes" Molecules 29, no. 17: 4081. https://doi.org/10.3390/molecules29174081

Article Metrics

Back to TopTop