Next Article in Journal
Suzuki–Miyaura Reaction in the Presence of N-Acetylcysteamine Thioesters Enables Rapid Synthesis of Biomimetic Polyketide Thioester Surrogates for Biosynthetic Studies
Next Article in Special Issue
Catalytic Performances of Co/TiO2 Catalysts in the Oxidative Dehydrogenation of Ethane to Ethylene: Effect of CoTiO3 and Co2TiO4 Phase Formation
Previous Article in Journal
Reactivity of Rare-Earth Oxides in Anhydrous Imidazolium Acetate Ionic Liquids
Previous Article in Special Issue
Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Au Single Metal Atom for Carbon Dioxide Reduction Reaction

1
Departament de Química, Universitat Autònoma de Barcelona, 08193 Cerdanyola del Vallès, Catalonia, Spain
2
Institut de Química Computacional i Catàlisi and Departament de Química, Universitat de Girona, c/Ma Aurèlia Capmany 69, 17003 Girona, Catalonia, Spain
*
Authors to whom correspondence should be addressed.
Chemistry 2023, 5(2), 1395-1406; https://doi.org/10.3390/chemistry5020095
Submission received: 4 March 2023 / Revised: 7 May 2023 / Accepted: 16 May 2023 / Published: 5 June 2023

Abstract

:
CO2 is the gas that contributes the most to the greenhouse effect and, therefore, to global warming. One of the greatest challenges facing humanity is the reduction of the concentration of CO2 in the air. Here, we analyze the possible use of Au1@g-C3N4 electrocatalyst to transform CO2 into added-value products. We use density functional theory (DFT) to determine the reaction Gibbs energies for eight electron–proton transfer reaction paths of the electrochemical carbon dioxide reduction reaction (CO2RR) using a single Au atom supported on 2D carbon nitride support. Our simulations classify the Au1@g-C3N4 electrocatalysts as “beyond CO” since their formation is energetically favored, although their strong binding with a Au single atom does not allow the desorption process. DFT calculations revealed that the lowest energy pathway is CO2 (g) → COOH* → CO* → HCO* → HCOH* → CH2OH* → CH2* → CH3* → CH4 (g), where the first hydrogenation of CO to HCO is predicted as the rate-limiting step of the reaction with slightly lower potential than predicted for Cu electrodes, the most effective catalysts for CO2RR. Methane is predicted to be the main reaction product after eight proton–electron transfers (CO2 + 8 H+ + 8e → CH4 + 2H2O). The generation of formaldehyde is discarded due to the large formation energy of the adsorbed moiety and the production of methanol is slightly less favorable than methane formation. Our computational study helps to identify suitable electrocatalysts for CO2RR by reducing the amount of metal and using stable and low-cost supports.

1. Introduction

One of the most challenging and relevant problems that our society must face to improve the quality of life of future generations is climate change, with the emission of carbon dioxide (CO2) considered the major contributor [1,2]. The use of renewable energy resources can drastically reduce these emissions, although nowadays approximately 80% of global energy consumption consists of fossil fuels [3]. Mitigation strategies like capture [4,5] and sequestration of CO2 [6,7], or even functionalization towards other chemicals [8,9], especially cycloaddition of CO2 to epoxides [10,11,12,13], have been pursued, although the conversion of CO2 to valuable fuels is more promising [3,14,15]. This could promote the use of desirable waste as feedstock for added-value compounds for the industry [16,17,18].
From the point of view of heterogeneous catalysis, the research community is making progress in creating smaller and dispersed metal particles supported on innocent and non-innocent supports based on metal oxides [19,20,21,22,23,24,25] and metal carbides [26,27,28,29], with Cu as the highest active metal particle for CO2 hydrogenation to methanol and Ni as the most effective metal for CO2 methanation [30].
In recent years, significant progress has been made to find active and selective catalysts for electrochemical CO2 reduction reaction (CO2RR) [31], where noble metals such as Pd [32], Ag [33,34], and Au [35,36], and transition metals such as Fe and Co enhance CO formation as the main product [37]. However, Cu is the rising star in this group of metal catalysts, since it is the only transition metal able to reduce CO2 into alcohols and hydrocarbons with acceptable Faradaic efficiencies [31,38,39,40,41,42,43,44,45,46]. A way to drastically reduce the cost of the catalysts and at the same time increase their catalytic activity and selectivity is to use single metal atoms. Over the past decade, single-atom catalysis (SAC) has become one of the most widely researched methods for the synthesis of novel low-cost and effective catalysts [47]. It is proposed that these single metal atoms have the same principle as homogeneous catalysts, improving the selectivity with respect to metal-supported nanoparticles, although the local coordination around the single atom site plays an important role [48]. In fact, SAC can be treated as the step following supported homogeneous catalysts, since after the removal of the ligand, it is stabilized as SAC [49]. The preparation of SACs and the support choice is not trivial, since it can promote clustering [50,51] and absorption inside the catalysts [52].
Carbon-based materials such as graphene have high surface areas and superior electrical conductivity and are normally employed as SAC supports to boost oxygen reduction reaction [53], oxygen evolution reaction [54], hydrogen evolution reaction [55], and CO2RR [56], among others, improving electrochemical activities. Graphitic carbon nitride (g-C3N4), with a graphene-like framework, contains periodic heptazine units connected via tertiary amines. The presence of nitrogen atoms in this graphene structure provides rich electron lone pairs to capture metal ions in the ligands. In fact, these electron pairs promote layer corrugation due to their electronic repulsions, although the corrugation phenomena enhance the generation of active sites for metal deposition [57]. 2D g-C3N4 is an excellent candidate for electrocatalytic CO2 reduction, acting as a non-innocent support for single metal atom catalysts—mainly Cu, Pd, and Pt [58,59,60]. Experimental studies using Cu1@g-C3N4 as a catalyst for CO2RR proved the stability of Cu single atom since the formation of Cu nanoparticles was not observed [59]. Moreover, it was found that formic acid and H2 were the main products. From a theoretical point of view, the CO2RR has been studied using Cu and Au atoms supported on polymeric carbon nitride (CN) [61] and several transition metal atoms embedded on C2N [62]. Focusing on polymeric CN, authors [61] highlighted the selectivity of Cu single atom to produce CH4; however, neither the competitive hydrogen evolution reaction (HER) nor the poor selectivity of Au single atom was addressed. On the other hand, the recent work by Dobrota et al. reported, by means of density functional simulations, the stability of several single metal atoms on similar carbon nitride supports, showing that Cu, Ag, and Au are stable but have lower stability compared with other transition metals [63]. Despite this, Cu1@g-C3N4 and Au1@g-C3N4 have been synthesized [64].
In this work, we want to explore the capability of a single Au atom supported on g-C3N4 for CO2RR. The single Au atom deposition on g-C3N4 was experimentally confirmed [64] but its catalytic activity has not been elucidated. This paper explores the reaction mechanism of the eight electron–proton transfer processes by means of periodic density functional calculations (DFT) applying the computational hydrogen electrode approach (CHE) [65].

2. Results

First, it is important to clarify the relevance of the monolayer corrugation. Initially, the optimized monolayer with and without the single metal atom has a flat structure. The deposition of Au single atom did not modify the planarity of the monolayer. However, the adsorption of all the moieties involved in the different pathways of CO2RR and HER process promoted surface corrugation (see Figure 1). This corrugation process is not unexpected and has been previously observed for g-C3N4 slab [52] and g-C3N4 monolayers [66,67]. According to Azofra et al. [57], corrugation minimizes the electronic repulsion of the nitrogen lone pairs located in their structural holes. From an energy point of view, the corrugated surface is around 1.2–2 eV lower in energy than the planar, including or not the transition metal single atom. Therefore, the global minimum and the reference of the monolayer’s energy is the corrugated structure, while the planar is a metastable structure that becomes corrugated after the adsorption of reactants. With respect to single-atom deposition, Au anchoring is favored by 1.91 eV. This result, together with the experimental characterization of Au1@g-C3N4 [58], highlights its stability.
Figure 2 shows the Gibbs energy profile, considering the lowest energy pathways, while Figure 3 illustrates the sketches of all the adsorbed species of CO2RR reaction on Au1@g-C3N4 and their relative Gibbs energies with respect to CO2 and the C3N4 surface in the gas phase. In a previous study, we elucidated the interaction and activation of CO2 molecules on Au1@g-C3N4, obtaining a binding energy of −0.22 eV, and maintaining the linear conformation of the molecule [61]. This was in contrast to other studies where CO2 was not bonded to Au single atom on 2D-carbon nitrides structure [68]. CO2 bending facilitates the proton-coupled electron transfer, although Cu single atom on g-C3N4 was found to be active both experimentally [59] and computationally [61] for the two proton-coupled electron transfers maintaining the (almost) linear structure. Going into detail about each proton-electron transfer, the first proton–coupled electron transfer is energetically favored, independent of the formed species, with COOH (2a, see Figure 3 for labels) being slightly lower in energy than HCOO (2b) by 0.33 eV. The second proton–electron transfer has two competitive pathways involving the ejection of one water molecule and the formation of adsorbed CO on top of the Au single atom or the generation of formic acid. Our simulations predict that the formation of adsorbed CO is favored by 0.83 eV, clearly ruling out the formation of HCOOH. However, our predictions show that CO would not be detected as a product since its desorption energy is very large (1.58 eV). Therefore, the Au1@g-C3N4 system can be classified into the “beyond CO” group of catalysts. The third proton–electron transfer, which implies the hydrogenation of CO, is the most energetically demanding step of the reaction since it has the largest thermodynamic barrier. This predicted barrier, at 0 V, gives the limiting potential value. It coincides with the reaction mechanism explored with Cu [38,43], where a limiting potential of −0.74 V is required. For the Au1@g-C3N4 system, the limiting potential is −0.69 eV, slightly lower than the most efficient catalyst for CO2RR. It is important to mention that CO* hydrogenates toward HCO* species (4b) because the formation of COH* is 2.00 eV higher in energy (4a). The HCO* is hydrogenated to HCOH* moiety (5b) during the fourth proton–electron transfer. The thermodynamic energy barrier for this step is only 0.08 eV, whereas the formation of the CH2O moiety (5c) comes with an energetic cost of 1.15 eV. The possible formation and desorption of one water molecule (5a) is not competitive since it implies an energy barrier larger than 3.5 eV. DFT simulations show that two moieties are close in energy for the fifth proton-coupled electron transfer. The hydrogenation of HCOH (5b) to CH2OH (6b) and CH3O (6c) is energetically favored, being that the formation of CH2OH is lower in energy by 0.23 eV. A priori, this small difference makes both hydrogenation processes feasible, although one should consider that the generation of CH3O from HCOH would involve, apart from the formation of the new C-H bond, cleavage of the O-H bond, which difficult its generation staring from HCOH species. With respect to water molecule formation (6a), this step is endergonic by more than 2 eV. The sixth proton–electron transfer is one of the most relevant steps because it may involve the generation of methanol and methane. According to our simulations, the hydrogenation of CH2OH (6b) and/or CH3O (6c) to CH4 (g) is endergonic by more than 1.4 eV in the best cases (7c and 7d). Thus, the conversion of CO2 with six protons to CH4 is not energetically feasible. In this case, the ejection of one water molecule and the formation of CH2* moiety (7a) is predicted as the lowest energy step with an energy barrier of 0.65 eV, followed by the production of methanol, which is 0.19 eV higher in energy (this will be discussed later). Following the lowest energy route, the next step involves the hydrogenation of CH2* (7a) to CH3* (8a). This process is exergonic and the previous step of CH4 formation at the eighth electron–proton transfer. The last step is endergonic by 0.66 eV, very close to the rate-limiting step. Thus, according to our simulations, methane is the main product of CO2RR when using Au1@g-C3N4 as a catalyst and applying a limiting voltage of −0.69 V that corresponds to CO hydrogenation to HCO. CO formation is predicted to be exergonic, i.e., without the requirement of a potential, although its strong binding with the Au atom is detrimental to its desorption. It is important to mention that CH4 production is possible after eight electron–proton transfers. The same product can be generated after six electron–proton transfers starting from CH2OH species, although our results show that the formation of one water molecule and its corresponding desorption is a much more favored process (by 1.00 eV).
A limit potential of −0.84 V is predicted for methanol production (Figure 4a). As illustrated in Figure 4, the hydrogenation of the CH2OH moiety (6b) to methanol (7b) is possible when the limiting potential is increased, making it the rate limiting step in the reaction. This process is competitive with respect to the ejection of one water molecule. However, the process would be feasible if CH3O (6c) is formed during the fifth proton–electron transfer. This moiety is only 0.25 eV larger in energy than CH2OH and its generation can promote methanol production with a limit potential of −0.69 V, the same value required to hydrogenate CO to HCO* (Figure 4b). Nevertheless, as has been mentioned before, the precursor of both moieties is HCOH (5b), which is clearly the lowest energy product of the fourth hydrogenation process. Thus, the formation of CH3O is not only a simple hydrogenation process but involves cleavage of the O-H bond and subsequent migration of the proton to the C atom. Furthermore, increasing the limiting potential will still favor ejection of the water molecule and formation of CH2* species (7a) bonded to Au single atom. For these reasons, the generation of methanol using Au1@g-C3N4 as an electrocatalyst is difficult, even if the limiting potential is increased. With respect to the formation of CH2O (g) (5c*), this process is endergonic and 0.55 eV higher than the formation of HCOH (5b). In this case, the major difficulty is the generation of the adsorbed CH2O* species that implies an energy cost larger than 1 eV, despite the fact that its desorption is favored in energy. Therefore, the generation of formaldehyde is excluded.
Finally, it is important to compare these results with the competitive reaction, the proton evolution to H2. In this case, the binding energy of H to the Au atom is −0.92 eV [63]. Therefore, a very large overpotential is required for H2 production, which favors the CO2RR on Au1@g-C3N4. Nevertheless, the large H–Au interaction can poison the catalytic active site, although the binding energy of COOH/HCOO moieties, the products of the first electron–proton transfer, and CO have higher binding energies than H. Moreover, Au is not the preferred site for H adsorption since the interaction with monolayer carbon atoms is higher. Thus, the monolayer would be covered by protons before binding to Au single atom.

3. Conclusions

In this work, DFT simulations have been carried out to explore the capability of single Au atom supported on g-C3N4 for CO2RR using the CHE approach. The lowest energy pathway is CO2 (g) → COOH* → CO* → HCO* → HCOH* → CH2OH* → CH2* → CH3* → CH4 (g), where the step that marks the limiting potential is the CO hydrogenation to HCO, the third electron–proton transfer. A limiting potential of −0.69 V is required to overcome this thermodynamic barrier; this is lower than the potential required for Cu metal, the most efficient electrocatalyst for CO2RR. Our simulations predict CH4 generation, where eight proton–electron transfers are required (CO2 + 8H+ + 8e → CH4 + 2H2O). The formation of CH4 using six protons (CO2 + 6H+ + 6e → CH4 + H2O + O*) is not feasible since the ejection of a second water molecule is preferred (CH2OH* + H+ + e → CH2* + H2O). In fact, this is the key step of the catalyst’s selectivity, since this water generation is slightly favored over methanol production by 0.18 eV. The high stability of the CH2OH* moiety with respect to CH3O* is detrimental to methanol production, where a limiting potential of −0.84 V is required.
In summary, applying the CHE approach, Au1@g-C3N4 is predicted as a promising electrocatalyst for CO2RR, avoiding H2 production due to the large overpotential required. This work encourages research on single metal atom properties for catalysis as well as the role of density functional simulations in order to shed light on the adsorption mode, coordination environment, and stability of single atom catalysts.

4. Computational Details and Models

The geometry of the bare monolayer, gas phase species, and the combination of both have been obtained from total energy minimization with the energy obtained from periodic DFT-based calculations carried out using the Vienna Ab initio Simulation Package (VASP) [69]. Exchange-correlation effects have been accounted for using generalized gradient approximation employing the Perdew-Burke-Ernzerhof (PBE) exchange-correlation potential [70], including the semi-empirical method of Grimme (D3) to describe the dispersion effects [71]. The valence electron wavefunctions were expanded onto a plane-wave basis set. The effect of the core electrons is considered through the projector augmented wave (PAW) method of Blöchl [72] as implemented by Kresse and Joubert [73]. The Brillouin zone was sampled by a 5 × 5 × 1 grid of k-points within the Monkhorst-Pack scheme [74]. The atomic positions were relaxed until the forces were smaller than 0.01 eV Å−1. The threshold for electronic relaxation was less than 10−5 eV.
The (001) monolayer of g-C3N4 is used as model (Cmcm space group), the most favorable available on materials project database [75]. The monolayer contains 24 C and 32 N atoms since the cell is expanded two times in the x and y axes. The Au deposition occurs between the Nitrogen atoms, which stabilize the system. The corrugation of the surface monolayer occurs after adsorption of the reactants [64]. Thus, to obtain the minimum energy value of the surface, the corrugated monolayers were computed without reactants, obtaining almost identical corrugated geometries degenerate in energy. The binding energy of single metal atom was calculated following Equation (1),
E ads = E Au @ C 3 N 4 ( E C 3 N 4 + E Au )
where E Au @ C 3 N 4 is the DFT energy of the monolayer including the anchored Au atom, E C 3 N 4 is the energy of the corrugated g-C3N4 monolayer, and E Au is the energy of one Au atom in its ground state.
In this work, we have evaluated the eight possible electron–proton transfer processes, considering all the possible reaction pathways:
CO2 (g) + H+ + e → COOH*
CO2 (g) + H+ + e → HCOO*
COOH*/HCOO* + H+ + e → CO* + H2O (g)
CO* → CO (g) + *
COOH*/HCOO* + H+ + e → HCOOH (g)
CO* + H+ + e → COH*
CO* + H+ + e → HCO*
HCO*/COH* + H+ + e → C* + H2O (g)
HCO*/COH* + H+ + e → HCOH*
HCO*/COH* + H+ + e → CH2O*
CH2O* → CH2O (g)
C*+ H+ + e → CH*
HCOH*/CH2O* + H+ + e → CH* + H2O (g)
HCOH*/CH2O* + H+ + e → CH2OH*
HCOH*/CH2O* + H+ + e → CH3O*
CH* + H+ + e → CH2*
CH2OH*/CH3O* + H+ + e → CH2* + H2O (g)
CH2OH*/CH3O* + H+ + e → CH3OH (g)
CH2OH*/CH3O* + H+ + e → CH4 (g) + O*
CH2* + H+ + e → CH3*
O* + H+ + e → OH*
CH3* + H+ + e → CH4 (g)
OH* + H+ + e → H2O (g)
To compute the reaction energy, we have employed the computational hydrogen electrode (CHE) approach proposed by Norskov and coworkers [65]. In this protocol, the energy of H+ + e can be computed as half of the Gibbs energy of the H2 molecule at 0 V vs. SHE. Therefore, the reaction energy of each step can be calculated by computing the binding energies of adsorbed moieties and the initial and final products (CO2, CO, CH2O, CH3OH, and CH4) in gas phase. The vibrational frequencies have also been determined through the construction and diagonalization of the Hessian matrix, constructed by independent displacements of atoms by 0.03 Å, and it has been used to compute the zero-point energy and the entropic effects considering adsorbate moieties. Accounting for the DFT energy, entropy, and ZPE, one can evaluate the Gibbs energies of the species involved in the reaction network. The Gibbs energies can be directly related to the electrode potential by following Equation (2) as an example case for reaction 4b:
ΔG4b = ΔGHCO* − 1/2 ΔGH2 − ΔGCO* + eUSHE
where e is the charge of the transferred electron and U is the voltage. The limiting potential required is the electrode potential at which all the reaction steps are exergonic, without thermodynamic barrier. On the other hand, solvation effects may be relevant to stabilize some reaction intermediates owing to the H-bonding effect of water molecules [62,76], although in our previous calculations, including solvents had no remarkable differences (0.01–0.05 eV) [66] and thus all the calculations in the present study are performed without including solvent effects.

Author Contributions

Investigation, A.V.-L. and S.P.-P.; writing, review and editing, S.P.-P., M.S. and A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministerio de Ciencia e Innovación (projects PID2021-127423NB-I00 to A.P. and PID2020-113711GB-I00 to M.S.) and the Generalitat de Catalunya (project 2021SGR0623 and ICREA Academia prize 2019 to A.P.). A.V.-L. is grateful for funding from the pre-doctoral fellowship (PRE2019-089647). S.P.-P. appreciates the economic support of the Marie Curie fellowship (H2020-MSCA-IF-2020-101020330).

Data Availability Statement

Data are deposited in the Universitat de Girona repository.

Acknowledgments

A.P. is a Serra Húnter Fellow and thanks the ICREA Academia Prize 2019. The article is in honor of Avelino Corma for the excellent contributions in the field of heterogeneous catalysis.

Conflicts of Interest

There are no conflict of interest to declare.

References

  1. Lim, L. How to make the most of carbon dioxide. Nature 2015, 526, 628–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Intergovernmental Panel on Climate Change. Climate Change 2013—The Physical Science Basis, 1st ed.; Cambridge University Press: Cambridge, UK, 2014. [Google Scholar]
  3. Luque-Urrutia, J.A.; Ortiz-García, T.; Solà, M.; Poater, A. Green Energy by Hydrogen Production from Water Splitting, Water Oxidation Catalysis and Acceptorless Dehydrogenative Coupling. Inorganics 2023, 11, 88. [Google Scholar] [CrossRef]
  4. Poater, J.; Gimferrer, M.; Poater, A. Covalent and Ionic Capacity of MOFs To Sorb Small Gas Molecules. Inorg. Chem. 2018, 57, 6981–6990. [Google Scholar] [CrossRef] [PubMed]
  5. Rajjak Shaikh, A.; Posada-Pérez, S.; Brotons-Rufes, A.; Pajski, J.J.; Kumar, G.; Mateen, A.; Poater, A.; Solà, M.; Chawla, M.; Cavallo, L. Selective absorption of H2S and CO2 by azole based protic ionic liquids: A combined Density Functional Theory and Molecular Dynamics study. J. Mol. Liq. 2022, 367, 120558. [Google Scholar] [CrossRef]
  6. Arayachukiat, S.; Yingcharoen, P.; Vummaleti, S.V.C.; Cavallo, L.; Poater, A.; D’Elia, V. Cycloaddition of CO2 to challenging N-tosyl aziridines using a halogen-free niobium complex: Catalytic activity and mechanistic insights. Mol. Catal. 2017, 443, 280–285. [Google Scholar] [CrossRef]
  7. Vummaleti, S.V.C.; Nolan, S.P.; Cavallo, L.; Talarico, G.; Poater, A. Mechanism of CO2 fixation by Ir–X Bonds (X = OH, OR, N, C). Eur. J. Inorg. Chem. 2015, 2015, 4653–4657. [Google Scholar] [CrossRef]
  8. Coufourier, S.; Gaignard-Gaillard, Q.; Lohier, J.-F.; Poater, A.; Gaillard, S.; Renaud, J.-L. Hydrogenation of CO2, Hydrogenocarbonate, and Carbonate to Formate in Water using Phosphine Free Bifunctional Iron Complexes. ACS Catal. 2020, 10, 2108–2116. [Google Scholar] [CrossRef]
  9. Darensbourg, D.J.; Holtcamp, M.W. Catalysts for the reactions of epoxides and carbon dioxide. Coord. Chem. Rev. 1996, 27, 155–174. [Google Scholar] [CrossRef]
  10. Vidal-López, A.; Posada-Pérez, S.; Solà, M.; D’Elia, V.; Poater, A. The Importance of the Bite Angle of Metal(III) Salen Catalysts in the Sequestration of CO2 with Epoxides in Mild Conditions. Green Chem. Eng. 2022, 3, 180–187. [Google Scholar] [CrossRef]
  11. Aomchad, V.; Del Gobbo, S.; Yingcharoen, P.; Poater, A.; D’Elia, V. Exploring the potential of Group III salen complexes for the conversion of CO2 under ambient conditions. Catal. Today 2021, 375, 324–334. [Google Scholar] [CrossRef]
  12. Natongchai, W.; Luque-Urrutia, J.A.; Phungpanya, C.; Solà, M.; D’Elia, V.; Poater, A.; Zipse, H. Cycloaddition of CO2 to epoxides by highly nucleophilic 4-aminopyridines: Establishing a relationship between carbon basicity and catalytic performance by experimental and DFT investigations. Org. Chem. Front. 2021, 8, 613–627. [Google Scholar] [CrossRef]
  13. Sodpiban, O.; Del Gobbo, S.; Barman, S.; Aomchad, V.; Kidkhunthod, P.; Ould-Chikh, S.; Poater, A.; D’Elia, V.; Basset, J.-M. Synthesis of Well-defined Yttrium-based Lewis Acids by Capture of a Reaction Intermediate and Catalytic Application for cycloaddition of CO2 to Epoxides Under Atmospheric Pressure. Catal. Sci. Technol. 2019, 9, 6152–6165. [Google Scholar] [CrossRef]
  14. Bobadilla, L.F.; Azancot, L.; Luque-Álvarez, L.A.; Torres-Sempere, G.; González-Castaño, M.; Pastor-Pérez, L.; Yu, J.; Ramírez-Reina, T.; Ivanova, S.; Centeno, M.A.; et al. Development of Power-to-X Catalytic Processes for CO2 Valorisation: From the Molecular Level to the Reactor Architecture. Chemistry 2022, 4, 1250–1280. [Google Scholar] [CrossRef]
  15. Schäppi, R.; Rutz, D.; Dähler, F.; Muroyama, A.; Haueter, P.; Lilliestam, J.; Patt, A.; Furler, P.; Aomchad, V.; Del Gobbo, S.; et al. Drop-in fuels from sunlight and air. Nature 2022, 601, 63–68. [Google Scholar] [CrossRef]
  16. Preti, D.; Resta, C.; Squarcialupi, S.; Fachinetti, G. Carbon dioxide hydrogenation to formic acid by using a heterogeneous gold catalyst. Angew. Chem. Int. Ed. 2011, 50, 12551–12554. [Google Scholar] [CrossRef]
  17. Wang, Y.; Darensbourg, D.J. Carbon dioxide-based functional polycarbonates: Metal catalyzed copolymerization of CO2 and epoxides. Coord. Chem. Rev. 2018, 372, 85–100. [Google Scholar] [CrossRef]
  18. Pomelli, C.S.; Tomasi, J.; Solà, M. Theoretical Study on the Thermodynamics of the Elimination of Formic Acid in the Last Step of the Hydrogenation of CO2 Catalyzed by Rhodium Complexes in the Gas Phase and Supercritical CO2. Organometallics 1998, 17, 3164–3168. [Google Scholar] [CrossRef]
  19. Italiano, C.; Llorca, J.; Pino, L.; Ferraro, M.; Antonucci, V.; Vita, A. CO and CO2 methanation over Ni catalysts supported on CeO2, Al2O3 and Y2O3 oxides. Appl. Catal. B Environ. 2020, 264, 118494. [Google Scholar] [CrossRef]
  20. Lam, E.; Noh, G.; Larmier, K.; Safonova, O.V.; Copéret, C. CO2 Hydrogenation on Cu-Catalysts Generated from ZnII Single-Sites: Enhanced CH3OH Selectivity Compared to Cu/ZnO/Al2O3. J. Catal. 2021, 394, 266–272. [Google Scholar] [CrossRef]
  21. Song, X.; Yang, C.; Li, X.; Wang, Z.; Pei, C.; Zhao, Z.J.; Gong, J. On the Role of Hydroxyl Groups on Cu/Al2O3 in CO2 Hydrogenation. ACS Catal. 2022, 12, 14162–14172. [Google Scholar] [CrossRef]
  22. Falbo, L.; Visconti, C.G.; Lietti, L.; Szanyi, J. The effect of CO on CO2 methanation over Ru/Al2O3 catalysts: A combined steady-state reactivity and transient DRIFT spectroscopy study. Appl. Catal. B Environ. 2019, 256, 117791. [Google Scholar] [CrossRef]
  23. Yan, Y.; Wong, R.J.; Ma, Z.; Donat, F.; Xi, S.; Saqline, S.; Fan, Q.; Du, Y.; Borgna, A.; He, Q.; et al. CO2 hydrogenation to methanol on tungsten-doped Cu/CeO2 catalysts. Appl. Catal. B Environ. 2022, 306, 121098. [Google Scholar] [CrossRef]
  24. Ye, R.-P.; Liao, L.; Reina, T.R.; Liu, J.; Chevella, D.; Jin, Y.; Fan, M.; Liu, J. Engineering Ni/SiO2 catalysts for enhanced CO2 methanation. Fuel 2021, 285, 119151. [Google Scholar] [CrossRef]
  25. Gonell, F.; Puga, A.V.; Julián-López, B.; García, H.; Corma, A. Copper-doped titania photocatalysts for simultaneous reduction of CO2 and production of H2 from aqueous sulfide. Appl. Catal. B Environ. 2016, 180, 263–270. [Google Scholar] [CrossRef] [Green Version]
  26. Posada-Pérez, S.; Ramirez, P.J.; Gutierrez, R.A.; Stacchiola, D.J.; Viñes, F.; Liu, P.; Illas, F.; Rodriguez, J.A. The conversion of CO2 to methanol on orthorhombic β-Mo2C and Cu/β-Mo2C catalysts: Mechanism for admetal induced change in the selectivity and activity. Catal. Sci. Technol. 2016, 6, 6766–6777. [Google Scholar] [CrossRef]
  27. Dixit, M.; Peng, X.; Porosoff, M.D.; Willauer, H.D.; Mpourmpakis, G. Elucidating the role of oxygen coverage in CO2 reduction on Mo2C. Catal. Sci. Technol. 2017, 7, 5521–5529. [Google Scholar] [CrossRef]
  28. Posada-Pérez, S.; Ramírez, P.J.; Evans, J.; Viñes, F.; Liu, P.; Illas, F.; Rodriguez, J.A. Highly active Au/δ-MoC and Cu/δ-MoC catalysts for the conversion of CO2: The metal/C ratio as a key factor defining activity, selectivity, and stability. J. Am. Chem. Soc. 2016, 138, 8269–8278. [Google Scholar] [CrossRef] [Green Version]
  29. Rodriguez, J.A.; Ramírez, P.J.; Gutierrez, R.A. Highly Active Pt/MoC and Pt/TiC Catalysts for the Low-Temperature Water-Gas Shift Reaction: Effects of the Carbide Metal/Carbon Ratio on the Catalyst Performance. Catal. Today 2017, 289, 47–52. [Google Scholar] [CrossRef]
  30. Posada-Pérez, S.; Solà, M.; Poater, A. Carbon dioxide conversion on supported metal nanoparticles: A brief review. Catalysts 2023, 13, 305. [Google Scholar] [CrossRef]
  31. Todorova, T.K.; Schreiber, M.W.; Fontcave, M. Mechanistic understanding of CO2 reduction reaction (CO2RR) toward multicarbon products by heterogeneous copper-based catalysts. ACS Catal. 2020, 10, 1754–1768. [Google Scholar] [CrossRef]
  32. Jiang, T.-W.; Zhou, Y.-W.; Ma, X.-Y.; Qin, X.; Li, H.; Ding, C.; Jiang, B.; Jiang, K.; Cai, W.-B. Spectrometric study of electrochemical CO2 reduction on Pd and Pd-B electrodes. ACS Catal. 2021, 11, 840–848. [Google Scholar] [CrossRef]
  33. Mistry, H.; Choi, Y.W.; Bagger, A.; Scholten, F.; Bonifacio, C.S.; Sinev, I.; Divins, N.J.; Zegkinoglou, I.; Jeon, H.S.; Kisslinger, K.; et al. Enhanced carbon dioxide electroreduction to carbon monoxide over defect-reach plasma-activated silver catalysts. Angew. Chem. Int. Ed. 2017, 56, 11394–11398. [Google Scholar] [CrossRef] [Green Version]
  34. Clark, E.L.; Ringe, S.; Tang, M.; Walton, A.; Hahn, C.; Jaramillo, T.F.; Chan, K.; Bell, A.T. Influence of atomic surface structure on the activity of Ag for the electrochemical reduction of CO2 to CO. ACS Catal. 2019, 9, 4006–4014. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, Y.; Li, C.W.; Kanan, M.W. Aqueous CO2 reduction at very low overpotential on oxide-derived Au nanoparticles. J. Am. Chem. Soc. 2012, 134, 19969–19972. [Google Scholar] [CrossRef]
  36. Mistry, H.; Reske, R.; Zeng, Z.; Zhao, Z.-J.; Greeley, J.; Strasser, P.; Roldan-Cuenya, B. Exceptional size-dependent activity enhancement in the electroreduction of CO2 over Au nanoparticles. J. Am. Chem. Soc. 2014, 136, 16473–16476. [Google Scholar] [CrossRef]
  37. Bonin, J.; Maurin, A.; Robert, M. Molecular catalysis of the electrochemical and photochemical reduction of CO2 with Fe and Co metal based complexes. Recent advances. Coord. Chem. Rev. 2017, 334, 184–198. [Google Scholar] [CrossRef]
  38. Rendón-Calle, A.; Builes, S.; Calle-Vallejo, F. How symmetry factors cause potential- and facet-dependent pathway shifts during CO2 reduction to CH4 on Cu electrodes. Curr. Opin. Electrochem. 2018, 9, 158–165. [Google Scholar] [CrossRef]
  39. Diaz-Morales, O.; Ledezma-Yanez, I.; Koper, M.T.M.; Calle-Vallejo, F. Guidelines for the Rational Design of Ni-Based Double Hydroxide Electrocatalysts for the Oxygen Evolution Reaction. ACS Catal. 2015, 9, 5380–5387. [Google Scholar] [CrossRef]
  40. Rendón-Calle, A.; Low, Q.H.; Hong, S.H.L.; Builes, S.; Yao, B.S.; Calle-Vallejo, F. A brief review of the computational modeling of CO2 electroreduction on Cu electrodes. Appl. Catal. 2021, 285, 119776. [Google Scholar] [CrossRef]
  41. Reske, R.; Mistry, H.; Behafarid, F.; Roldan-Cuenya, B.; Strasser, P. Particle size effects in the catalytic electroreduction of CO2 on Cu nanoparticles. J. Am. Chem. Soc. 2014, 136, 6978–6986. [Google Scholar] [CrossRef]
  42. Woldu, A.R.; Huang, Z.; Zhao, P.; Hu, L.; Astruc, D. Electrochemical CO2 reduction (CO2RR) to multi-carbon products over copper-based catalysts. Coord. Chem. Rev. 2022, 454, 214340. [Google Scholar] [CrossRef]
  43. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Norskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  44. Nie, X.; Luo, W.; Janik, M.J.; Asthagiri, A. Reaction mechanisms of CO2 electrochemical reduction on Cu(1 1 1) determined with density functional theory. J. Catal. 2014, 312, 108–122. [Google Scholar] [CrossRef]
  45. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar] [CrossRef]
  46. Goodpaster, J.D.; Bell, A.T.; Head-Gordon, M. Identification of Possible Pathways for C–C Bond Formation during Electrochemical Reduction of CO2: New Theoretical Insights from an Improved Electrochemical Model. J. Phys. Chem. Lett. 2016, 7, 1471–1477. [Google Scholar] [CrossRef]
  47. Datye, A.K.; Guo, H. Single atom catalysis poised to transition from an academic curiosity to an industrially relevant technology. Nat. Commun. 2021, 12, 895. [Google Scholar] [CrossRef]
  48. Liu, L.; Corma, A. Metal catalysts for heterogeneous catalysis: From single atoms to nanoclusters and nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [Green Version]
  49. Basset, J.M.; Baudouin, A.; Bayard, F.; Candy, J.-P.; Copéret, C.; De Mallmann, A.; Godard, G.; Kuntz, E.; Lefebvre, F.; Lucas, C.; et al. Modern Surface Organometallic Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009; pp. 23–73. [Google Scholar]
  50. Rivera-Cárcamo, C.; Serp, P. Single atom catalysts on carbon-based materials. ChemCatChem 2018, 10, 5058–5091. [Google Scholar] [CrossRef]
  51. Xu, Y.-T.; Xie, M.-Y.; Zhong, H.; Cao, Y. In situ clustering of single-atom copper precatalysts in a metal-organic framework for efficient electrocatalytic nitrate-to-ammonia reduction. ACS Catal. 2022, 12, 8698–8706. [Google Scholar] [CrossRef]
  52. Jurado, L.; Esvan, J.; Luque-Álvarez, L.A.; Bobadilla, L.F.; Odriozola, J.A.; Posada-Pérez, S.; Poater, A.; Comas-Vives, A.; Axet, M.R. Highly dispersed Rh single atoms over graphitic carbon nitride as a robust catalyst for the hydroformylation reaction. Catal. Sci. Technol. 2023, 13, 1425–1436. [Google Scholar] [CrossRef]
  53. Wang, Y.; Li, Q.; Zhang, L.c.; Wu, Y.; Chen, H.; Li, T.; Xu, M.; Bao, S.-J. A gel-limiting strategy for large-scale fabrication of Fe-N-C single-atom ORR catalysts. J. Mater. Chem. A 2021, 9, 7137–7142. [Google Scholar] [CrossRef]
  54. Ying, Y.; Fan, K.; Luo, X.; Qiao, J.; Huang, H. Unravelling the origin of bifunctional OER/ORR activity for single-atom catalysts supported on C2N by DFT and machine learning. J. Mater. Chem. A 2021, 9, 16860–16867. [Google Scholar] [CrossRef]
  55. Zeng, L.; Dai, C.; Liu, B.; Xue, C. Oxygen-assisted stabilization of single-atom Au during photocatalytic hydrogen evolution. J. Mater. Chem. A 2019, 7, 24217–24221. [Google Scholar] [CrossRef]
  56. Li, M.; Wang, H.; Luo, W.; Sherrell, P.C.; Chen, J.; Yang, J. Heterogeneous single-atom catalysts for electrochemical CO2 reduction reaction. Adv. Mater. 2020, 32, e2001848. [Google Scholar] [CrossRef]
  57. Azofra, L.M.; MacFarlane, D.R.; Sun, C. A DFT study of planar vs. corrugated graphene-like carbon nitride (g-C3N4) and its role in the catalytic performance of CO2 conversion. Phys. Chem. Chem. Phys. 2016, 18, 18507–18514. [Google Scholar] [CrossRef]
  58. Gao, G.; Jiao, Y.; Waclawik, E.; Du, A. Single Atom (Pd/Pt) Supported on Graphitic Carbon Nitride as an Efficient Photocatalyst for Visible-Light Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2016, 138, 6292–6297. [Google Scholar] [CrossRef] [Green Version]
  59. Cometto, C.; Ugolotti, A.; Grazietti, E.; Moretto, A.; Bottaro, G.; Armelao, L.; Di Valentin, C.; Calvillo, L.; Granozzi, G. Copper single-atoms embedded in 2D graphitic carbon nitride for the CO2 reduction. npj 2D Mater. Appl. 2021, 5, 63. [Google Scholar] [CrossRef]
  60. Peng, Y.; Lu, B.; Chen, S. Carbon-Supported Single Atom Catalysts for Electrochemical Energy Conversion and Storage. Adv. Mater. 2018, 30, 1801995. [Google Scholar] [CrossRef]
  61. Li, J.; Yan, O.; Li, K.; You, J.; Wang, H.; Cui, W.; Cen, W.; Chu, Y.; Dong, F. Cu supported on polymeric carbon nitride for selective CO2 reduction into CH4: A combined kinetic and thermodynamics investigation. J. Mater. Chem. A 2019, 7, 17014–17021. [Google Scholar] [CrossRef]
  62. Liu, H.; Huang, Q.; An, W.; Wang, Y.; Men, Y.; Liu, S. Dual-atom active sites embedded in two-dimensional C2N for efficient CO2 electroreduction: A computational study. J. Energy Chem. 2021, 61, 507–516. [Google Scholar] [CrossRef]
  63. Dobrota, A.S.; Skorodumova, N.V.; Mentus, S.V.; Pašti, I.A. Surface Pourbaix plots of M@N4-graphene single-atom electrocatalysts from density functional theory thermodynamic modeling. Electrochim. Acta 2022, 412, 140155. [Google Scholar] [CrossRef]
  64. Yu, J.; Chen, C.; Zhang, Q.; Lin, J.; Yang, X.; Gu, L.; Zhang, H.; Liu, Z.; Wang, Y.; Zhang, S.; et al. Au Atoms Anchored on Amorphous C3N4 for Single-Site Raman Enhancement. J. Am. Chem. Soc. 2022, 144, 21908–21915. [Google Scholar] [CrossRef] [PubMed]
  65. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  66. Posada-Pérez, S.; Vidal-López, A.; Solà, M.; Poater, A. 2D carbon nitride as a support with single Cu, Ag, and Au atoms for carbon dioxide reduction reaction. Phys. Chem. Chem. Phys. 2023, 25, 8574–8582. [Google Scholar] [CrossRef]
  67. Zhao, X.; Wang, L.; Pei, Y. Single metal atom catalysts supported on g-C3N4 for formic acid dehydrogenation: A combining density functional theory and machine learning study. J. Phys. Chem. C 2021, 125, 22513–22521. [Google Scholar] [CrossRef]
  68. Cui, X.; An, W.; Liu, X.; Wang, H.; Men, Y.; Wang, J. C2N-graphene supported single-atom catalysts for CO2 electrochemical reduction reaction: Mechanistic insight and catalysts screening. Nanoscale 2018, 10, 15262–15272. [Google Scholar] [CrossRef]
  69. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B Condens. Matter Mater. Phys. 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  70. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  71. 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, 154104–154119. [Google Scholar] [CrossRef] [Green Version]
  72. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  73. Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  74. Wisesa, P.; McGill, K.; Mueller, T. Efficient generation of generalized Monkhorst-Pack grids through the use of informatics. Phys. Rev. B 2016, 93, 155109. [Google Scholar] [CrossRef] [Green Version]
  75. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef] [Green Version]
  76. Feng, Y.; An, W.; Wang, Z.; Wang, Y.; Men, Y.; Du, Y. Electrochemical CO2 reduction reaction on M@Cu(211) bimetallic single-atom surface alloys: Mechanism, kinetics, and catalyst screening. ACS Sustain. Chem. Eng. 2020, 8, 210–222. [Google Scholar] [CrossRef]
Figure 1. Sketches of the top (up) and lateral (down) corrugated Au1/g-C3N4 monolayer.
Figure 1. Sketches of the top (up) and lateral (down) corrugated Au1/g-C3N4 monolayer.
Chemistry 05 00095 g001
Figure 2. The lowest energy profile of CO2RR calculated using periodic DFT simulations using Au1/g-C3N4 at 0V and a limiting potential of −0.69 V corresponding to the energy barrier of CO hydrogenation to HCO.
Figure 2. The lowest energy profile of CO2RR calculated using periodic DFT simulations using Au1/g-C3N4 at 0V and a limiting potential of −0.69 V corresponding to the energy barrier of CO hydrogenation to HCO.
Chemistry 05 00095 g002
Figure 3. Full CO2RR reaction scheme using Au1/g-C3N4 at 0V. The Gibbs energy values in parentheses (eV) correspond to the relative reaction energies with respect to CO2 (g) and the Au1/g-C3N4 surface. Energies in bold and black arrows mark the lowest energy path.
Figure 3. Full CO2RR reaction scheme using Au1/g-C3N4 at 0V. The Gibbs energy values in parentheses (eV) correspond to the relative reaction energies with respect to CO2 (g) and the Au1/g-C3N4 surface. Energies in bold and black arrows mark the lowest energy path.
Chemistry 05 00095 g003
Figure 4. Reaction scheme of methanol as a product of CO2RR using Au1/g-C3N4 (a) considering CH2OH intermediate and (b) considering CH3O intermediate.
Figure 4. Reaction scheme of methanol as a product of CO2RR using Au1/g-C3N4 (a) considering CH2OH intermediate and (b) considering CH3O intermediate.
Chemistry 05 00095 g004
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

Vidal-López, A.; Posada-Pérez, S.; Solà, M.; Poater, A. Au Single Metal Atom for Carbon Dioxide Reduction Reaction. Chemistry 2023, 5, 1395-1406. https://doi.org/10.3390/chemistry5020095

AMA Style

Vidal-López A, Posada-Pérez S, Solà M, Poater A. Au Single Metal Atom for Carbon Dioxide Reduction Reaction. Chemistry. 2023; 5(2):1395-1406. https://doi.org/10.3390/chemistry5020095

Chicago/Turabian Style

Vidal-López, Anna, Sergio Posada-Pérez, Miquel Solà, and Albert Poater. 2023. "Au Single Metal Atom for Carbon Dioxide Reduction Reaction" Chemistry 5, no. 2: 1395-1406. https://doi.org/10.3390/chemistry5020095

Article Metrics

Back to TopTop