Next Article in Journal
Laser-Induced Synthesis of Electrocatalytically Active Ag, Pt, and AgPt/Polyaniline Nanocomposites for Hydrogen Evolution Reactions
Next Article in Special Issue
Catalyst Design: Counter Anion Effect on Ni Nanocatalysts Anchored on Hollow Carbon Spheres
Previous Article in Journal
Long-Term Antifogging Coating Based on Black Phosphorus Hybrid Super-Hydrophilic Polymer Hetero-Network
Previous Article in Special Issue
Recent Application of Core-Shell Nanostructured Catalysts for CO2 Thermocatalytic Conversion Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Icosahedral Copper-Based Bimetallic Clusters for the Selective Electrocatalytic CO2 Conversion to One Carbon Products

by
Azeem Ghulam Nabi
1,2,3,4,*,
Aman-ur-Rehman
2,5,6,
Akhtar Hussain
4,
Gregory A. Chass
1,7,8 and
Devis Di Tommaso
1,*
1
Department of Chemistry, School of Physical and Chemical Sciences, Queen Mary University of London, Mile End Road, London E1 4NS, UK
2
Department of Physics and Applied Mathematics, Pakistan Institute of Engineering and Applied Sciences, Nilore, Islamabad 45650, Pakistan
3
Department of Physics, University of Gujrat, Jalalpur Jattan Road, Gujrat 50700, Pakistan
4
Theoretical Physics Division, Pakistan Institute of Nuclear Science& Technology (PINSTECH), Nilore, Islamabad 45650, Pakistan
5
Department of Nuclear Engineering, Pakistan Institute of Engineering & Applied Sciences, Nilore, Islamabad 45650, Pakistan
6
Center for Mathematical Sciences, Pakistan Institute of Engineering & Applied Sciences, Nilore, Islamabad 45650, Pakistan
7
Department of Chemistry, McMaster University, Hamilton, ON L8S 4L8, Canada
8
Faculty of Land and Food Systems, The University of British Columbia, Vancouver, BC V6T1Z4, Canada
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 87; https://doi.org/10.3390/nano13010087
Submission received: 16 November 2022 / Revised: 19 December 2022 / Accepted: 20 December 2022 / Published: 24 December 2022
(This article belongs to the Special Issue Nanocatalysts for Methanation Reaction)

Abstract

:
Electrochemical CO2 reduction reactions can lead to high value-added chemical and materials production while helping decrease anthropogenic CO2 emissions. Copper metal clusters can reduce CO2 to more than thirty different hydrocarbons and oxygenates yet they lack the required selectivity. We present a computational characterization of the role of nano-structuring and alloying in Cu-based catalysts on the activity and selectivity of CO2 reduction to generate the following one-carbon products: carbon monoxide (CO), formic acid (HCOOH), formaldehyde (H2C=O), methanol (CH3OH) and methane (CH4). The structures and energetics were determined for the adsorption, activation, and conversion of CO2 on monometallic and bimetallic (decorated and core@shell) 55-atom Cu-based clusters. The dopant metals considered were Ag, Cd, Pd, Pt, and Zn, located at different coordination sites. The relative binding strength of the intermediates were used to identify the optimal catalyst for the selective CO2 conversion to one-carbon products. It was discovered that single atom Cd or Zn doping is optimal for the conversion of CO2 to CO. The core@shell models with Ag, Pd and Pt provided higher selectivity for formic acid and formaldehyde. The Cu-Pt and Cu-Pd showed lowest overpotential for methane formation.

1. Introduction

The rising carbon dioxide (CO2) level and overall concentrations in the atmosphere due to fossil fuel combustion, a major cause of global warming, pose a serious threat to humankind [1]. One of the most promising solutions to mitigating this risk is via the chemical conversion of gaseous CO2 into value-added chemicals and materials [2]. The electrochemical CO2 reduction reaction (eCO2RR) has emerged as a potential strategy for converting CO2 because if coupled with electricity from renewable sources (wind, solar, or hydropower plants), the eCO2RR could achieve a carbon-neutral energy cycle [3,4]. The main challenges in eCO2RR lie in the activation of competitive CO2-minimizing pathways such as the hydrogen evolution reaction (HER, H+ + e → ½ H2) [5,6] and the conversion of CO2 to a specific product with good selectivity; this given the marginal difference in the electrochemical potentials of CO2 reduction into different products [3]. For example, the change for transforming CO2 to ethylene is −0.34 V while the CO2 to methanol transformation is −0.38 V, relative to the standard hydrogen electrode (SHE) [7].
Catalysts can facilitate favorable pathways to reduce the overall energy requirements of eCO2RR. Due to their ability to activate CO2, initial research focused on noble metal-based catalysts (Pt, Rh, Ir) [8,9,10,11,12], yet their scarcity and cost have retarded development. Hence, earth-abundant and active metal-based catalysts have been become requisite to develop sustainable solutions to CO2 transformation, when all aspects are considered (i.e., fundamental chemistry/physics, technological, economical). Copper (Cu) emerges as the best candidate for eCO2RR, being the only metal surface that reduces CO2 to more than thirty hydrocarbons and oxygenates [13], yet lacks the required selectivity [14,15,16]. Relevant studies dedicated to improving selectivity and hindering the HER have investigated the adsorption/desorption mechanism on single crystal Cu electrodes to demonstrate the role of surface morphology [17]. It was discovered that Cu crystal facets with high index planes such as Cu(711) are more selective in the production of valuable two-carbon (C2) products, such as ethylene and ethanol. This with respect to the dominant Cu(111) surface [18], while stepped Cu surfaces such as the (211) facet more easily produce one-carbon (C1) hydrocarbons [19]. Computational studies also revealed that the higher activity of polycrystalline Cu nanoparticles is due to the presence of stepped facets, such as (110) [20], (211) [21] and Cu(321) [22]. These stepped surfaces occur in metal clusters [18,22,23,24], where both the number of uncoordinated sites at the corners and edges [25] and the surface-to-volume ratio of nanoparticles are higher than those on copper surfaces, which may lead to improved catalytic properties towards eCO2RR [13].
Another strategy to improve the activity and selectivity of Cu electrodes is metal (M) doping [14]. Bimetallic catalysts often show better catalytic performance than the corresponding elemental metal ones due to synergic effects between the two metallic centers [26]. The dopant provides reaction sites with varied electronic properties and modulates those of the host (Cu), influencing the adsorption strength of the eCO2RR intermediates. Experimental studies have also revealed that low doping concentrations facilitate the formation of C1 products [27,28]. In particular, metal dopants such as Ag [28,29], Cd [30], Pd [28,31], Pt [32], and Zn [28,33] in Cu-M catalysts show efficiency towards C1/C2 products.
Consequently, in Cu-based catalysts, the nanoscale structuring and cooperative metal-metal coupling could enhance CO2 activation and selectivity, leading to specific product formation. In this regard, quantum mechanical modelling has provided insights into the structure, stability and catalytic properties of CuM clusters, while also demonstrating that an appropriate proportion of metal atoms influences the CO2 activation and selectivity towards the desired reaction. Alvarez-Garcia et al. investigated the binding and dissociation of CO2 on four-atom bimetallic CunPd4−n (n = 0–4) clusters employing density functional theory (DFT) calculations [30], resolving the ideal composition for adsorption energy and facile barriers to activation barrier was found in Cu3Pd; in agreement with the Pd/(Pd + Cu) atomic ratioing reported experimentally [34,35]. Investigation on the effect of substituting Cu with Zr on CO2 adsorption for the four-atom Cu4 cluster [36] revealed that the energy barriers for the direct dissociation of the CO2 molecule to CO + O decreased significantly for bimetallic CuZr clusters, with respect to pure Cu4. Our recent computational work on small tetrahedral CuSn clusters revealed the Cu2Sn2 system to suppress the competitive HER, while being highly selective towards the electrochemical CO2 → CO conversion [37]. Xing et al. considered bimetallic PdnCum (m + n = 15, with n > m) clusters wherein Pd10Cu5 showed the highest catalytic activity, particularly towards the CO2 → COOH hydrogenation step [31]. Li et al. considered (Cu)n clusters with n = 8, 20, 38 (even numbers) and n = 13, 55 (odd numbers) to investigate the reactivity at the high density corner and edge sites and found the icosahedral Cu55 to provide the lowest energy pathway to the CO intermediate and the ensuing C2 ethylene product [25].
Computational characterizations of clusters in the size range of 10 ≤ n ≤ 55 showed that (Cu)n adopted the icosahedral structure [38] derived from 13- and 55-atom icosahedra, built by adding or removing atoms. In addition, a comparison of icosahedral and cuboctahedral (n = 55, 147 and 309) clusters confirmed the icosahedral copper clusters to be more stable. Experimental verification of the formation of copper clusters using microemulsion techniques revealed Cu55 to be one of the most abundant clusters followed by Cu13, Cu147 and Cu309 [39]. According to a recent DFT investigation, Cu55 exhibits highly degenerate states [40]; a direct outcome of its icosahedral symmetry. Therefore, study on nanoclusters such as the highly symmetric 55-atom icosahedral structures would give a deeper understanding than stepped surfaces. This has been attributed to their larger surface-to-volume ratio and higher proportion of coordinatively unsaturated surface atoms (corner or edge) in comparison to bulk materials, resulting in a narrowing of the d-band, an upward shift of the band’s energy, and consequently, a stronger adsorption of the reaction intermediates [41]. Investigation on the adsorption of CO2 on icosahedral 55-atom Cu-based bimetallic clusters [42] found that for the Cu55−xZrx systems (x = 1–12), the formation of the CO2-activated state (linear to bent transition and elongation of C–O bonds) was endothermic on the pure copper cluster but barrierless and exothermic on the Zr-decorated system. Similarly, DFT calculations of Cu55−xZrx systems (x = 0, 12, 13, 42, 43 and 55) with a core@shell and decorated distribution of Cu and Ni atoms showed the presence of Ni on the clusters was crucial to the activation of CO2 [43]. Although previous computational studies of icosahedral Cu-based bimetallic nanocatalysts considered the adsorption, activation and gas-phase dissociation of CO2, in the context of eCO2RR, the focus should be on the concerted proton-electron transfer (CPET) steps [44].
Here, we present a computational investigation based on DFT calculations of the effect of nano-structuring and alloying in Cu-based catalysts on the activity and selectivity of the eCO2RR. Starting from the icosahedral Cu55 structure, we generated Cu54M1, Cu43M12 and Cu30M25 decorated architectures and Cu13M42 core@shell models (M = Ag, Cd, Pd, Pt, and Zn) (Figure 1), with the metals located at three different coordination sites (6, 8 and 12). We provide a thorough analysis of the structural, thermodynamic and electronic properties of these nanoclusters and their ability to activate CO2. The computational hydrogen electrode (CHE) model [45] was then applied to compute the mechanism of eCO2RR to the C1 products carbon monoxide (CO), formic acid (HCOOH), formaldehyde (CH2O), methane (CH4) and methanol (CH3OH). We have focused our attention to C1 products because a recent techno-economic assessment of low-temperature CO2 electrolysis shows the production costs of C1 products such as HCOOH and CO are competitive to conventional processes compared to C2 products such as ethylene and ethanol, which production has substantially higher costs [46] We compare the free energy profiles for the electrocatalytic CO2 conversion to these C1 products to the competitive HER. The relative binding strength of the intermediates involved is used to identify catalysts for the selective CO2 conversion. For comparison purposes, calculations of the eCO2RR and HER were also conducted on the (100), (110), (111) and (211) facets of pure copper.

2. Computational Methods

2.1. Atomistic Models of Clusters and Surfaces

The icosahedral (Ih) 55-atom monometallic Cu cluster was generated using the ab initio random structure searching (AIRSS) code [47]. The decorated Cu54M clusters were then generated by replacing one surface Cu with a dopant metal atom M, where M = Ag, Cd, Pd, Pt and Zn. As shown in Figure 1a, there are three possible coordination sites: CN6 is the edge site, CN8 is the corner site and CN12 is the center of the nanocluster. The Cu43M12 model in Figure 1b was generated by replacing 12 Cu atoms with M located at CN6. The Cu25M30 model in Figure 1c was generated by replacing 12 Cu atoms with M located at CN8. The Cu13M42 core@shell model in Figure 1d was generated by replacing all 13 surface Cu atoms with M. We also considered four-layer (3 × 3) slab models of Cu(100), Cu(110), Cu(111) and Cu(211) [20] with the (100), (110) and (111) being the dominant surfaces of copper. The Cu(211) facet was considered because of its good selectivity towards C1 formation. This was linked to the Cu(211) morphology characterized by step-edge sites with a coordination number equal to 7 (CN7) [48]. Here, we have also compared the catalytic conversion of CO2 to C1 chemicals on Cu(211) to that on 55-atoms icosahedral Cu-M nanoclusters with M located at CN6 and CN8.

2.2. Density Functional Theory Calculations

Calculations of energies and structures were conducted at the spin-polarized DFT level using the “Vienna ab initio simulation package” (VASP Software GmbH, version 6.3.1, Vienna, Austria) [49] using the following computational settings: the Perdew–Burke–Ernzerhof (PBE) exchange correlation functional with the Grimme’s-D3 dispersion correction; a plane-wave basis set within the framework of the projector augmented wave method with a kinetic energy cutoff (Ecut) set to 400 eV; a single k-point (1 × 1 × 1) for the nanoclusters and a (5 × 5 × 1) k-point mesh for the surface model to sample the Brillouin zone of the simulation supercell; a 0.18 eV width for the smearing. Energies, zero-point energies, and entropies of H2(g), CO2(g) and CO(g), and H2O used to compute the free energy corrections are reported in Supplementary Information (Table S1).

2.3. Free Energy Calculations

Following the computational hydrogen electrode (CHE) method proposed by Nørskov and co-workers [45], the Gibbs free energy of each step involved in the eCO2RR to C1 products was computed using the following equation:
Δ G = Δ S + Δ E ZPE T Δ S + Δ G solv + Δ G U
where ΔE is the reaction energy; ΔEZPE is the change in zero-point energy; ΔS is the change in entropy and T is the temperature of the reaction (300 K). We determined the latter two quantities within the harmonic approximation by taking the vibrational frequencies of adsorbates and molecules calculated with DFT. The solvation effects to compute the solvation free energy term ΔGsolv were included using VASPsol [50]. ΔGU is the free energy correction introduced by the difference of the electrode potential. For reactions involving a concerted proton–electron transfer (CPET) step, the ΔGU term can be computed by applying the formula:
Δ G U = n e U
where n is the number of electrons transferred, e is the electron charge and U is the applied electrode potential. The limiting potential (UL) and the overpotential (η) are important factors for evaluating the catalytic activity. The limiting potential is given by the formula:
U L = Δ G max / n e
where ∆Gmax is the relative change of the Gibbs free energy of the rate-determining step. The overpotential (η) can be obtained by calculating the difference between the equilibrium potential (Ueq) and the limiting potential:
η = U eq U L
Thus, η represents the minimum applied potential required to facilitate the formation of relevant intermediates.

3. Results and Discussion

3.1. Stability, Structure, and Electronic Properties of the Icosahedral 55-Atom CuM Clusters

The segregation energy (SE) was used to determine the preference of the metal dopants (Ag, Cd, Pd, Pt and Zn) to be in the core or shell of Cu54M. The SE is defined as [51]:
S E = E [ Cu 54 M ( surface ) ] E [ Cu 54 M ( core ) ]
where E[Cu54M(surface)] and E[Cu54M(core)] are the electronic energies of the fully optimized Cu54M1 cluster obtained by replacing one Cu atom with a dopant metal at a surface (CN6 or CN8) and at the center of the cluster (CN12), respectively. In Figure 2, the SE values are negative for all Cu54M, which implies that the metal prefers to be at the surface of the cluster, consistent with DFT calculations of Cu54Zr [51]. The metal doping at the CN8 site is more stable than CN6, but because their values of SE were close, the adsorption and reduction of CO2 on both coordination sites.
To gain insights into the relative stability of pure and bimetallic 55-atom systems, we used the binding energy per atom (EB), defined as [52]:
E B = E ( Cu 55 x M x ) ( 55 x ) E ( Cu ) x E ( M ) 55
where E(Cu55−xMn) is the total energy of the most stable isomer of each Cu55−xMx cluster and E(Cu) and E(M) are the total energies of the Cu and Sn atoms, respectively. A higher negative value of EB indicates higher thermodynamic stability of the cluster. The calculated EB for pure Cu55 nanocluster is –2.99 eV, equal to the value obtained using all-electron triple-z quality DFT calculations [53]. Table 1 reports the calculated EB and other structural and electronic properties: the average interatomic bonding distance between nearest neighbors, the energy difference between the highest occupied molecular orbital (HOMO), the lowest unoccupied molecular orbital (LUMO) (ΔH−L), the Bader charge difference between the Cu and M atoms (∆QM), and the surface energy (γ). The surface energy was computed using the following equation [54]:
γ = E n a n o s p h e r e   N E b u l k 4 π R 2  
where Enanosphere is the energy of a cluster with N atoms (N = 55), Ebulk is the energy of the bulk material per one layer of cross-section and R is the radius a spherical incorporating the cluster.
A descriptor to analyze the global reactivity descriptor is the gap energy ΔH−L, which relates to the energy cost for an electron to jump from the HOMO to the LUMO orbital. Therefore, ΔH−L characterizes the chemical stability of the system, with a higher value corresponding to a more chemically stable (less reactive) cluster. The ΔH−L for pure Cu55 atom is 0.028 eV, consistent with the literature value of 0.03 eV [55]. In Table 1 and Figure S1b of Supplementary Information, the single-doped atom clusters Cu54Ag (CN6) and Cu54Zn (CN6) show small ΔH−L values of 1 × 10−4 and 5 × 10−4 eV, respectively. The single atom doped Cu54Ag (CN8) and Cu54Zn (CN8) show larger ΔH−L values of 4.9 × 10−2 and 4.1 × 10−2 eV. This shows that the coordination environment of the metal dopant affects the gap energy ΔH−L and, therefore, the reactivity of the cluster. Overall, the values of ΔH−L are between 1 × 10−4 and 1.7 × 10−1 eV. The highest value of ΔH−L, 1.7 × 10−1 eV, is found for Cu43Zn12. The charge distribution in CuM clusters depends on the doping metal. This will influence CO2 adsorption and subsequent CO2 reduction because the electron transfer occurs from the electron-rich metal to the C atom, which in CO2 is in its highest oxidation state. In the decorated clusters (Cu54M, Cu43M12, and Cu25M30), when M is Ag, Pd or Pt, charge is transferred from Cu to M (negative ∆QM), and when M is Cd or Zn charge is transferred from M to Cu (positive ∆QM). In the core@shell Cu13M42 architecture, when M = Ag, Pd and Pt, the core is positively charged because of the charge transfer from Cu to M, while the shell has a negative charge. Vice versa for Cu13M42 with M = Cd and Zn. The effect of atomic radii, covalent radii, van-der radii, and electronegativity difference (∆EN) on bond lengths and surface area are discussed and available in Supplementary Information (Table S2). Regarding the surface energy, all clusters have negative γ values signifying the stability of the clusters compared to the bulk.

3.2. Adsorption and Activation of CO2 on Cu and CuM Clusters

CO2 is a linear molecule with two equivalent C–O bonds (length = 1.12 Å).13 Before its dissociation, the first step in the catalytic conversion of CO2 is its adsorption on the catalyst surface. CO2 can maintain the geometric properties of gas-phase CO2 (physisorption) or become activated because of the charge transferred from the metal catalyst to the π* molecular orbitals of the CO2 molecule (chemisorption) resulting in the elongation of the C–O bonds and decrease in the O–C–O bond angle (linear to bent mode) [56]. Here, we have conducted a detailed characterization of the adsorption and activation of CO2 on the pure copper cluster Cu55 and the copper-metal clusters, Cu54M (CN6 and CN8), Cu43M12 (CN6), Cu25M30 (CN8) and Cu13M42 (core@shell), with M = Ag, Cd, Pd, Pt and Zn. These models can provide insights into the influence of surface chemistry on the activation of the CO2 molecule. The structures of CO2 on the CuM clusters are shown in Figure 3. The associated values of adsorption energies (Eads), bond angles, bond elongations and Bader charges of CO2 adsorbed on the CuM clusters are listed in Table 2. The adsorption energy was calculated as
E a d s = E C u M · · · C O 2 E C u M E C O 2
where the first term is the total energy of the CuM···CO2 system, and the second and third terms are the energies of the isolated cluster and CO2 molecules, respectively. CO2 is physisorbed on all Cu-Ag and core@shell clusters as indicated by the absence of significant deviations of the bond angle, bond elongation of adsorbed CO2 from the gas-phase values, and small charge transfer between Cu and M (QM ~0.04e). In Figure 3 and Table 2, η(Cu, C) and η(M, C) refer to configurations in which the C atom of the CO2 molecule is coordinated to the Cu and M atoms, respectively. In each chemisorbed state, there is a decrease in the O–C–O angle and an increase in charge transfer. In the single metal-doped systems, Cu54M, at the CN6 active site, the coordination state η(Cu, C) occurs for M = Cd and Zn and η(M, C) occurs for M = Pd and Pt. In single metal-doped clusters at CN8, the η(Cu, C) exists for M = Cd and η(M, C) is present for M = Pd and Pt. Both η(Cu, C) and η(Zn, C) exist for Cu54Zn on the CN8 active site. In the 12-atom doped Cu43M12 clusters, η(Cu, C) is present for M = Cd, Zn and η(M, C) exists in all systems except for Cu43Ag12. The 30-atom doped nano catalysts show the same trend as the 12-atom, except for the absence of η(Cd, C) in Cu43Cd12. The η(Cu, C) and η(M, C) coordination do not exist in the core@shell models because CO2 only physisorbs. In terms of adsorption energy, in the absence of η(M, C), the physisorption energy of CO2 always dominates. Similarly, in the absence of η(Cu, C), the chemisorption energy of η(M, C) configuration always dominates. When both η(Cu, C) and η(M, C) occur on a particular site, then again η(Cu, C) is the most stable coordination mode.

3.3. Mechanism of CO2 Reduction Reaction to C1 Products on Cu-M Clusters and Cu Surfaces

In this section, we present calculations of the mechanism of electrochemical CO2 reduction. Scheme 1 shows the pathways and intermediates for the formation of the following C1 products: CO, HCOOH, CH2O, CH4 and CH3OH. Depending on the atom coordinated to the catalyst, O or C, the first CPET step leads to two intermediates, *OCHO and *COOH. The second CPET will determine whether the 2e products HCOOH or CO is formed. Subsequent CPET will lead to 4e (CH2O), 6e (CH3OH) and 8e (CH4) C1 products. Compared to other catalytic reactions, the pathway of the eCO2RR is more complex because of the number of intermediates involved. According to Equation 1, the optimal reaction pathway is determined by the lowest free energy pathway at the applied potential U.

3.3.1. Electrocatalytic CO2 conversion to CO and HCOOH

We computed the free energy of reactions (ΔG) of the elementary steps to the CO2 conversion to HCOOH and CO on the following systems: icosahedral Cu55 cluster; decorated and core@shell Cu-M bimetallic clusters; Cu(100), Cu(110), Cu(111) and Cu(211) surfaces. In the context of the CHE model (see Equation 1), we define the potential limiting step (ΔGPLS) as the elementary reaction in the eCO2RR to CO or HCOOH (at U = 0 V) with the highest ΔG value (a high ΔGPLS corresponds to poor catalytic performance). The elementary steps leading to the formation of CO are: (i) CO2 adsorption (CO2 → *CO2, ΔG*CO2); (ii) CPET to convert *CO2 to C-coordinated formate (*CO2 + H+ + e → *COOH, ΔG*COOH); (iii) CPET to convert formate to adsorbed carbon monoxide (*COOH + H+ + e → *CO + H2O, ΔG*CO); (iv) the release, from the catalyst surface, of gas-phase CO (*CO → CO(g), ΔGCO). The structures of the optimized *COOH and *CO on all Cu and CuM systems are reported in Supplementary Information (Figures S2–S8).
Pure copper: CO2-to-CO conversion. The Gibbs free energy diagrams for the CO2-to-CO conversion on the monometallic 55-atom cluster and the (100), (110), (111) and (211) surfaces are reported in Figure 4a. There is a significant dependence of the stability of the *COOH and *CO intermediates on the surface morphology and coordination environment. The (211) facet has better catalytic performance (lower ΔGPLS) towards CO formation than any other surfaces but higher than Cu55. This cluster was, therefore, taken as the reference system to assess the performance of the bimetallic clusters. The competitive HER (H+ + e → ½ H2) in Figure 4b shows a similar morphology dependence: unfavourable on the (110) surface; highly favorable on the (100) surface; moderately favorable on the (211) surface and the Cu55 cluster.
Bimetallic clusters: CO2-to-CO conversion. The free energy diagrams for the eCO2RR to CO and the HER on CuM are reported in Figure 5. In the single metal doped clusters, Cu54M, the value of ΔGPLS depends on both the coordination site and nature of the metal. The ΔGPLS is lower when the reaction occurs on CN6 for M equal to Cd (0.16 eV), Pd (0.23 eV) and Pt (0.53 eV) compared to CN8, Cd (0.12 eV), Pd (0.42 eV) and Pt (0.78 eV). However, for Ag (0.27 eV) and Zn (0.18 eV), CN6 shows higher ΔGPLS than CN8, Ag (0.14 eV) and Zn (0.14 eV). Each intermediate shows strong chemisorption with a low ΔGPLS value and vice versa. For both CN6 and CN8 systems, HER is dominant (lower ΔGH) over eCO2RR because of the strong CO binding to the cluster, leading to a large ΔG*CO; the exception is M = Pt. When the number of metal dopants on the CuM cluster increases, Cu43M12, so does the value of ΔGPLS: Ag (0.24 eV), Pd (0.32 eV), Pt (0.86 eV) and Zn (0.19 eV). The exception is Cd (0.25 eV). The CO generation remains dominant over HER, except again for Pt, for the same reasons discussed for the single atom doped clusters. With further increase in the doping and change in surface chemistry in the Cu25M30 clusters, the HER becomes more favorable with a small value of |ΔGH| for Ag (0.10 eV), Cd (0.19 eV), Pd (0.92 eV) and Pt (0.76 eV) compared to the ΔGPLS of the eCO2RR of Ag (0.37 eV), Cd (0.29 eV), Pd (1.61 eV) and Pt (1.23 eV). At this doping concentration, only Zn, with ΔGPLS = 0.40 eV and ΔGH = 0.49 eV, favors eCO2RR over HER. All core@shell models are more active towards HER than eCO2RR: the ΔGH values of Ag (0.07 eV), Cd (0.45eV), Pd (−0.62 eV), Pt (−0.59 eV) and Zn(−0.27 eV) are lower than the ΔGPLS values of Ag (0.76 eV), Cd (0.91 eV), Pd (1.12 eV), Pt (1.09 eV), and Zn (0.58 eV).
Bimetallic clusters: CO2-to-HCOOH conversion. For HCOOH formation, the steps are the CPET to convert adsorbed *CO2 to O-coordinated formate (*CO2 + H+ + e → *OCOH, ΔG*OCOH) and the CPET to convert adsorbed formate to liquid phase formic acid (*OCHO + H+ + e → HCOOH(l), ΔGHCOOH). In Figure 5, the ΔGPLS value for HCOOH of Cu54M with CN6 for Ag (0.33 eV), Pd (0.33 eV) and Pt(0.37 eV) in the CuM cluster with CN8 for Ag(0.46 eV), Cd (0.46 eV), Pt(0.45 eV), Cu43M12 with Ag (0.39 eV), Pt (0.39 eV), Cu25M30 with Pd (0.19 eV), Pt (0.13 eV) and finally core@shell with Pd (0.33 eV), Pt (0.41 eV) are dominant over HER. We can explain this behaviour by considering the value of the d-band center (Table 2), which decreases for Ag, Cd and Zn with increasing doping concentration. The adsorption energy of the intermediates involved in the CO or HCOOH reaction pathway also decreases. Similarly, the higher position of the d-band center for Pd and Pt leads to an increase in the intermediate adsorption energy. Therefore, the core@shell model with low d-band center, Ag (–3.56 eV), Cd (–7.29 eV) and Zn (–6.23 eV), show poor catalytic performance, and Pd (–1.58 eV) and Pt (–1.92 eV) show good catalytic performance for HCOOH. Overall, the core@shell promotes the formation of HCOOH, and single metal-doped clusters show good catalytic performance for CO, except for Pt, which catalyzes HCOOH formation.

3.3.2. Electrocatalytic CO2 conversion to CH2O, CH3OH, and CH4

The free energy diagrams for the eCO2RR to formaldehyde (CH2O), methanol (CH3OH) and methane (CH4) on the CuM clusters are reported in Figure 6.
Bimetallic clusters: CH2O formation. After the eCO2RR reduction to *CO or *HCOOH, further CPET steps generates three distinct intermediates: *CHO, *COH or *OCH. Among them, *CHO is the easiest to generate, as illustrated by the free energy diagram for the formation of these species, where the lowest ∆GPLS is for COH. A subsequent CPET step leads to the formation of formaldehyde: *CHO + H+ + e → *OCH2 → * + CH2O(g). Due to the stronger O-affinity, CH2O prefers *OCHO than the *COOH route. CH2O generation shows lowest ΔGPLS values on Cu25Pd30 (0.19 eV) and core@shell Cu@Pd (0.33 eV). As the ΔGH values on Cu25Pd30 (0.92 eV) and core@shell Cu@Pd (0.62 eV) are higher than ΔGPLS values, the CO2 conversion to CH2O is dominant over HER on these clusters. CuPd is also favorable towards CH2O formation as the ΔGPLS values are higher than ΔGH on these clusters. The values of ΔGPLS are 0.45 eV for Cu54Pt (CN8), 0.49 eV for Cu25Pt30, and 0.44 eV for the Cu54Pt (CN8) and the core@shell Cu25Pt30 clusters. In comparison, the ΔGH values on Cu54Pt (CN8), Cu25Pt30 and Cu@Pt are 0.46 eV, 0.76 eV and 0.59 eV, respectively. The Gibbs free energy diagram of the eCO2RR to CH2O on these systems is given as Figure 6a–e
Bimetallic clusters: CH3OH formation. The formation of CH3OH involves five CPET steps. The first three reduction steps coincide to the eCO2RR to CH2O. The *CHO is reduced to *CHOH or *OCH2 depending on if the O or C atoms ate protonated. This leads to four possible routes to convert CO2 to CH3OH:
  • *COOH → *CO → *CHO → *OCH2→ *OCH3 → *OHCH3;
  • *OCHO → *OCH2O → *OCH2OH → *O + CH3OH → *OH → * + H2O;
  • *OCHO → *HCOOH → *CHO → *CHOH → *CH2OH → *OHCH3;
  • *OCHO → *HCOOH → *CHO → *OCH2 → *OHCH2 → *OHCH3.
Out of these four paths, our calculations predict the last one as the most suitable for CH3OH formation. Similar to CH2O, the Cu25Pd30 and core@shell Cu@Pd show the lowest ΔGPLS, 0.28 eV and 0.33 eV, respectively, and still, these reactions are dominant over HER.
Bimetallic clusters: CH4 formation. The main 8-electron product of eCO2RR is CH4, involving eight CPET transfer steps. It follows five different reaction pathways:
  • *CHO → *CHOH → *CH → *CH2 → *CH3 → * + CH4;
  • *CHO → *CHOH → *CH2OH → *CH2→ *CH3 → * + CH4;
  • *CHO → *OCH2 → *OHCH2 → *OHCH3 → *OH + CH4 → * + H2O;
  • *CHO → *OCH2 → *OCH3 → *OHCH3 → *OH + CH4 → * + H2O;
  • *CHO → *OCH2 → *OCH3 → *O + CH4 → *OH → * + H2O.
The free energy diagrams along these pathways are given in Figures S9 and S10 of Supplementary Information.
Competition between CH4 and H2 formation. HER is a competitive reaction in eCO2RR and can reduce the efficiency of the eCO2RR reaction leading to poor selectivity of the catalyst. To evaluate the selectivity of eCO2RR vs. HER, we have reported in Figure 7 the limiting potential difference ∆∆GPLS = ∆GPLS(eCO2RR) − ∆GH(HER) for the five reaction pathways leading to the formation of CH4 on the CuM clusters. The higher the (positive) value of ∆∆GPLS, the higher the selectivity for CH4. Only CuPd shows good catalytic performance towards CH4 formation. Through the reaction pathway (1), Cu25Pd30 has a positive ∆∆GPLS. The Cu43Ag12 and Cu43Cd12 clusters show a small negative ∆∆GPLS value. For pathways (2) and (3), the Cu25Pd30 and Cu@Pd show a positive ∆∆GPLS value and hence these two systems are selective towards CH4. Like pathway (1), the Cu43Ag12 and Cu43Cd12 show a very small ∆∆GPLS for (2) which makes them also good candidates for catalyzing CH4 formation. Finally, Cu13Cd42 shows a small ∆∆GPLS, which can be explained based on the d-band center and coordination environment: at the same doping concentration (1-atom) with CN6 and CN8, the 1-atom at CN6 show significantly low overpotential for all pathways leading to CH4 generation. Single-doped clusters with CN6 and CN8 have similar d-band center values (Table 2), only the coordination environment is different, which means that the CN environment has a significant impact on the catalytic performance. Furthermore, the d-band center value increases with an increase in doping concentration for Cu-Pd and Cu-Pt catalysts. Consequently, there is strong adsorption of the intermediates involved in the reaction (pathways 1 to 5). However, with M = Ag, Cd and Zn, the values of d-band centers decreases, which leads to weak adsorption of the intermediates and poor catalytic performance towards CH4 formation.

3.3.3. Selectivity

The overpotentials (η) to C1 products for all CuM systems, summarized in Figure 8, were computed from the equilibrium (Ueq) and limiting (UL) potentials (Figure 5 and Figure 6). The values of Ueq for CO, HCOOH, CH2O, CH3OH and CH4 are 0.12 V, 0.25 V, 0.07 V, 0.02 V and 0.17 V, respectively. For the Cu clusters doped with Ag, Cd and Zn, an increase in metal doping, particularly after 30-atom, leads to the HER becoming dominant over other C1 products. This behaviour is clearly noticeable for the single atom doped clusters, Cu54M with M = Ag, Cd and Zn, which shows higher η values for HER than the corresponding core@shell systems. The values of the overpotential also show that a single atom doped system supports either CO or HCOOH. Therefore, small doping does not promote CH3OH or CH4 formation. As metal doping increases, the Cu-Pd and Cu-Pt clusters show lower overpotential for CH2O, CH3OH and CH4. Only Cu25Pt30 and Cu13@Pd42 have the lowest overpotential for CH4.

4. Conclusions

In this work, the catalytic properties towards the electrochemical CO2 reduction reaction of a series of icosahedral 55-atom Cu-based clusters doped with Ag, Cd, Pd, Pt and Zn were investigated using density functional theory calculations. The adsorption and activation of CO2 on these clusters and all possible reaction paths that lead to the CO2 reduction to C1 products (CO, HCOOH, CH2O, CH3OH and CH4) were considered. Apart from the composition effects, the role of the coordination environment of the metal dopant on the catalytic performance of copper-based clusters was also investigated, with the results showing that nanoclusters with eight-coordinated metal dopants have better catalytic activity towards CO2 activation. Single-atom doping with Cd and Zn is the best candidate for the CO2-to-CO conversion, while core@shell with Ag, Pd and Pt is a good choice for formic acid or formaldehyde formation. The CuPt and CuPd systems show the lowest overpotential for methane formation. This work identifies the influence of size, metal coupling and metal coordination on CO2 activation and intermediate stability and, consequently, the structure-property relationship in Cu-based mono and bi-metallic clusters for the selective CO2 conversion to value-added C1 products.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13010087/s1, Table S1: The energies (E), zero-point energies (ZPE), and entropies (S) of H2(g), CO2(g) and CO(g), and H2O; Table S2: The atomic, covalent and Van der Waals radii, the electronegativity difference, electronic configuration, and calculated value of segregation energies (in eV) [57,58]; Figure S1: (a) The binding energy and (b) HOMO-LUMO (H-L) gap of Cu-M clusters with increasing doping concentration; Figure S2: The structure and adsorption energies (in eV) of COOH adsorbed on the Cu-M clusters; Figure S3: The structure and adsorption energies (in eV) of CO adsorbed on the Cu54M clusters with CN6 and CN8 nano-catalysts; Figure S4: The structure and adsorption energies (in eV) of CO on the Cu43M12 and Cu25M30 clusters; Figure S5: The structure and adsorption energies (in eV) of CO on the Cu43M12 and Cu25M30 clusters; Figure S6: The structure and adsorption energies of H adsorbed on the Cu43M12 and Cu25M30 clusters at the Top, Hollow and Bridge positions; Figure S7: The structure and adsorption energies of H adsorbed on the Cu43M12 and Cu25M30 clusters at Top, Hollow and Bridge positions; Figure S8: The structures and adsorption energies of H adsorbed on the core@shell clusters at Top, Hollow and Bridge positions; Figure S9. Gibbs free energy diagram for the CH4 formation on CuM clusters along pathways 1 to 5; Figure S10. Gibbs free energy diagram for CHO (a) and COH (b) formation on CuM clusters.

Author Contributions

Conceptualization of work: A.G.N. and D.D.T.; Conducting of experiments: A.G.N.; Computation: A.G.N.; Data analyses: A.G.N., A.-u.-R., A.H. and D.D.T.; Data dissemination & graphics: A.G.N., A.H. and D.D.T.; Writing of manuscript: A.G.N., A.-u.-R., G.A.C., D.D.T.; Project support: G.A.C., D.D.T. All authors have read and agreed to the published version of the manuscript.

Funding

A.G.N. acknowledges the Pakistan HEC-QMUL PhD Scholarships for funding. D.D.T. and G.A.C. acknowledged the ACT programme (Accelerating CCS Technologies, Horizon2020 Project No 299668), which funded the FUNMIN project. Financial contributions made from Department for Business, Energy & Industrial Strategy (BEIS) together with extra funding from NERC and EPSRC research councils, United Kingdom, ADEME (FR), MINECO-AEI (ES). We are grateful to the UK Materials and Molecular Modelling Hub for computational resources, which is partially funded by EPSRC (EP/P020194/1). We are grateful to the UK Materials and Molecular Modelling Hub for computational resources, which is partially funded by EPSRC (EP/P020194/1). Via our membership of the UK’s HEC Materials Chemistry Consortium, which is funded by EPSRC (EP/L000202), this work used the ARCHER UK National Supercomputing Service.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This research utilized Queen Mary’s Apocrita HPC facility, supported by QMUL Research-IT. http://doi.org/10.5281/zenodo.438045, accessed on 16 December 2022.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IPCC. Climate Change 2022: Mitigation of Climate Change; Working Group III Report; Cambridge University Press: Cambridge, UK, 2022; p. 1454. [Google Scholar]
  2. IPCC. Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems; Shukla, P.R., Skea, J., Calvo Buendia, E., Masson-Delmotte, V., Pörtner, H.-O., Roberts, D.C., Zhai, P., Slade, R., Connors, S., van Diemen, R., et al., Eds.; Cambridge University Press: Cambridge, UK, 2019. [Google Scholar]
  3. Bushuyev, O.S.; de Luna, P.; Dinh, C.T.; Tao, L.; Saur, G.; van de Lagemaat, J.; Kelley, S.O.; Sargent, E.H. What Should We Make with CO2 and How Can We Make It? Joule 2018, 2, 825–832. [Google Scholar] [CrossRef] [Green Version]
  4. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals. Adv. Sci. 2018, 5, 1700275. [Google Scholar] [CrossRef] [PubMed]
  5. Bagger, A.; Ju, W.; Varela, A.S.; Strasser, P.; Rossmeisl, J. Electrochemical CO2 Reduction: A Classification Problem. ChemPhysChem 2017, 18, 3266–3273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Peterson, A.A.; Nørskov, J.K. Activity descriptors for CO2 electroreduction to methane on transition-metal catalysts. J. Phys. Chem. Lett. 2012, 3, 251–258. [Google Scholar] [CrossRef]
  7. Wu, J.; Huang, Y.; Ye, W.; Li, Y. CO2 Reduction: From the Electrochemical to Photochemical Approach. Adv. Sci. 2017, 4, 1700194. [Google Scholar] [CrossRef]
  8. He, Z.; Qian, Q.; Ma, J.; Meng, Q.; Zhou, H.; Song, J.; Liu, Z.; Han, B. Water-Enhanced Synthesis of Higher Alcohols from CO2 Hydrogenation over a Pt/Co3O4 Catalyst under Milder Conditions. Angew. Chemie Int. Ed. 2016, 55, 737–741. [Google Scholar] [CrossRef]
  9. Zhang, X.; Li, X.; Zhang, D.; Su, N.Q.; Yang, W.; Everitt, H.O.; Liu, J. Product selectivity in plasmonic photocatalysis for carbon dioxide hydrogenation. Nat. Commun. 2017, 8, 14542. [Google Scholar] [CrossRef] [Green Version]
  10. Park, S.; Bézier, D.; Brookhart, M. An Efficient Iridium Catalyst for Reduction of Carbon Dioxide to Methane with Trialkylsilanes. J. Am. Chem. Soc. 2012, 134, 11404–11407. [Google Scholar] [CrossRef]
  11. Klinkova, A.; de Luna, P.; Dinh, C.-T.; Voznyy, O.; Larin, E.M.; Kumacheva, E.; Sargent, E.H. Rational Design of Efficient Palladium Catalysts for Electroreduction of Carbon Dioxide to Formate. ACS Catal. 2016, 6, 8115–8120. [Google Scholar] [CrossRef]
  12. Zhu, W.; Michalsky, R.; Metin, Ö.; Lv, H.; Guo, S.; Wright, C.J.; Sun, X.; Peterson, A.A.; Sun, S. Monodisperse Au Nanoparticles for Selective Electrocatalytic Reduction of CO2 to CO. J. Am. Chem. Soc. 2013, 135, 16833–16836. [Google Scholar] [CrossRef]
  13. Xie, H.; Wang, T.; Liang, J.; Li, Q.; Sun, S. Cu-based nanocatalysts for electrochemical reduction of CO2. Nano Today 2018, 21, 41–54. [Google Scholar] [CrossRef]
  14. Dickinson, H.L.A.; Symes, M.D. Recent progress in CO2 reduction using bimetallic electrodes containing copper. Electrochem. Commun. 2022, 135, 107212. [Google Scholar] [CrossRef]
  15. Kitchin, J.R.; Nørskov, J.K.; Barteau, M.A.; Chen, J.G. Role of Strain and Ligand Effects in the Modification of the Electronic and Chemical Properties of Bimetallic Surfaces. Phys. Rev. Lett. 2004, 93, 156801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sinfelt, J.H.; Carter, J.L.; Yates, D.J.C. Catalytic hydrogenolysis and dehydrogenation over copper-nickel alloys. J. Catal. 1972, 24, 283–296. [Google Scholar] [CrossRef]
  17. Shen, S.; He, J.; Peng, X.; Xi, W.; Zhang, L.; Xi, D.; Wang, L.; Liu, X.; Luo, J. Stepped surface-rich copper fiber felt as an efficient electrocatalyst for the CO2RR to formate. J. Mater. Chem. A 2018, 6, 18960–18966. [Google Scholar] [CrossRef]
  18. Hori, Y.; Takahashi, I.; Koga, O.; Hoshi, N. Selective formation of C2 compounds from electrochemical reduction of CO2 at a series of copper single crystal electrodes. J. Phys. Chem. B 2002, 106, 15–17. [Google Scholar] [CrossRef]
  19. Sreejyothi, P.; Mandal, S.K. From CO2 activation to catalytic reduction: A metal-free approach. Chem. Sci. 2020, 11, 10571–10593. [Google Scholar]
  20. Higham, M.D.; Quesne, M.G.; Catlow, C.R.A. Mechanism of CO2 conversion to methanol over Cu(110) and Cu(100) surfaces. Dalt. Trans. 2020, 49, 8478–8497. [Google Scholar] [CrossRef]
  21. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F.; et al. The Mechanism of CO and CO2 Hydrogenation to Methanol over Cu-Based Catalysts. ChemCatChem 2015, 7, 1105–1111. [Google Scholar] [CrossRef] [Green Version]
  22. Fajín, J.L.C.; Cordeiro, M.N.D.S.; Illas, F.; Gomes, J.R.B. Influence of step sites in the molecular mechanism of the water gas shift reaction catalyzed by copper. J. Catal. 2009, 268, 131–141. [Google Scholar] [CrossRef]
  23. Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.; et al. The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science 2012, 759, 893–898. [Google Scholar] [CrossRef]
  24. Hori, Y.; Takahashi, I.; Koga, O.; Hoshi, N. Electrochemical reduction of carbon dioxide at various series of copper single crystal electrodes. J. Mol. Catal. A Chem. 2003, 199, 39–47. [Google Scholar] [CrossRef]
  25. Li, Q.; Zhang, Y.; Shi, L.; Wu, M.; Ouyang, Y.; Wang, J. Dynamic structure change of Cu nanoparticles on carbon supports for CO2 electro-reduction toward multicarbon products. InfoMat 2021, 3, 1285–1294. [Google Scholar] [CrossRef]
  26. Steinhauer, B.; Kasireddy, M.R.; Radnik, J.; Martin, A. Development of Ni-Pd bimetallic catalysts for the utilization of carbon dioxide and methane by dry reforming. Appl. Catal. A Gen. 2009, 366, 333–341. [Google Scholar] [CrossRef]
  27. Wang, X.; Chen, Q.; Zhou, Y.; Li, H.; Fu, J.; Liu, M. Cu-based bimetallic catalysts for CO2 reduction reaction. Adv. Sens. Energy Mater. 2022, 1, 100023. [Google Scholar] [CrossRef]
  28. Zaza, L.; Rossi, K.; Buonsanti, R. Well-Defined Copper-Based Nanocatalysts for Selective Electrochemical Reduction of CO2 to C2Products. ACS Energy Lett. 2022, 7, 1284–1291. [Google Scholar] [CrossRef]
  29. Chen, C.; Li, Y.; Yu, S.; Louisia, S.; Jin, J.; Li, M.; Ross, M.B.; Yang, P. Cu-Ag Tandem Catalysts for High-Rate CO2 Electrolysis toward Multicarbons. Joule 2020, 4, 1688–1699. [Google Scholar] [CrossRef]
  30. Alvarez-Garcia, A.; Flórez, E.; Moreno, A.; Jimenez-Orozco, C. CO2 activation on small Cu-Ni and Cu-Pd bimetallic clusters. Mol. Catal. 2020, 484, 110733. [Google Scholar] [CrossRef]
  31. Xing, M.; Guo, L.; Hao, Z. Theoretical insight into the electrocatalytic reduction of CO2 with different metal ratios and reaction mechanisms on palladium–copper alloys. Dalt. Trans. 2019, 48, 1504–1515. [Google Scholar] [CrossRef]
  32. Han, L.; Liu, H.; Cui, P.; Peng, Z.; Zhang, S.; Yang, J. Alloy Cu3Pt nanoframes through the structure evolution in Cu-Pt nanoparticles with a core-shell construction. Sci. Rep. 2014, 4, 6414. [Google Scholar] [CrossRef] [Green Version]
  33. Jeon, H.S.; Timosnenko, J.; Scholten, F.; Sinev, I.; Herzog, A.; Haase, F.T.; Cuenya, B.R. Operando insight into the correlation between the structure and composition of CuZn nanoparticles and their selectivity for the electrochemical CO2 reduction. J. Am. Chem. Soc. 2019, 141, 19879–19887. [Google Scholar] [CrossRef] [PubMed]
  34. Feng, X.; Jiang, K.; Fan, S.; Kanan, M.W. Grain-Boundary-Dependent CO2 Electroreduction Activity. J. Am. Chem. Soc. 2015, 137, 4606–4609. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for selective CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2015, 170–171, 173–185. [Google Scholar] [CrossRef]
  36. Megha; Mondal, K.; Ghanty, T.K.; Banerjee, A. Adsorption and Activation of CO2 on Small-Sized Cu-Zr Bimetallic Clusters. J. Phys. Chem. A 2021, 125, 2558–2572. [Google Scholar] [CrossRef]
  37. Muthuperiyanayagam, A.; Azeem, G.N.; -ur-Rehman, A.; Di Tommaso, D. Adsorption, activation, and conversion of carbon dioxide on small copper-tin nanoclusters. ChemRxiv 2022. preprint. [Google Scholar] [CrossRef]
  38. Kabir, M.; Mookerjee, A.; Bhattacharya, A.K. Structure and stability of copper clusters: A tight-binding molecular dynamics study. Phys. Rev. A-At. Mol. Opt. Phys. 2004, 69, 043203. [Google Scholar] [CrossRef] [Green Version]
  39. Vázquez-Vázquez, C.; Bañobre-López, M.; Mitra, A.; López-Quintela, M.A.; Rivas, J. Synthesis of small atomic copper clusters in microemulsions. Langmuir 2009, 25, 8208–8216. [Google Scholar] [CrossRef]
  40. Häkkinen, H.; Moseler, M.; Kostko, O.; Morgner, N.; Hoffmann, M.A.; Issendorff, B.V. Symmetry and electronic structure of noble-metal nanoparticles and the role of relativity. Phys. Rev. Lett. 2004, 93, 093401. [Google Scholar] [CrossRef] [Green Version]
  41. Kleis, J.; Greeley, J.; Romero, N.A.; Morozov, V.A.; Falsig, H.; Larsen, A.H.; Lu, J.; Mortensen, J.J.; Dułak, M.; Thygesen, K.S.; et al. Finite Size Effects in Chemical Bonding: From Small Clusters to Solids. Catal. Letters 2011, 141, 1067–1071. [Google Scholar] [CrossRef]
  42. Austin, N.; Ye, J.; Mpourmpakis, G. CO2 activation on Cu-based Zr-decorated nanoparticles. Catal. Sci. Technol. 2017, 7, 2245–2251. [Google Scholar] [CrossRef]
  43. Austin, N.; Butina, B.; Mpourmpakis, G. CO2 activation on bimetallic CuNi nanoparticles. Prog. Nat. Sci. Mater. Int. 2016, 26, 487–492. [Google Scholar] [CrossRef]
  44. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  45. Rossmeisl, J.; Qu, Z.W.; Zhu, H.; Kroes, G.J.; Nørskov, J.K. Electrolysis of water on oxide surfaces. J. Electroanal. Chem. 2007, 607, 83–89. [Google Scholar] [CrossRef]
  46. Shin, H.; Hansen, K.U.; Jiao, F. Techno-economic assessment of low-temperature carbon dioxide electrolysis. Nat. Sustain. 2021, 4, 911–919. [Google Scholar] [CrossRef]
  47. Pickard, C.J.; Needs, R.J. Ab initio random structure searching. J. Phys. Condens. Matter. 2011, 23, 53201. [Google Scholar] [CrossRef] [Green Version]
  48. Xu, Y.; Li, F.; Xu, A.; Edwards, J.P.; Hung, S.F.; Gabardo, C.M.; O’Brien, C.P.; Liu, S.; Wang, X.; Li, Y.; et al. Low coordination number copper catalysts for electrochemical CO2 methanation in a membrane electrode assembly. Nat. Commun. 2021, 12, 4–10. [Google Scholar] [CrossRef]
  49. 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]
  50. Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T.A.; Hennig, R.G. Implicit solvation model for density-functional study of nanocrystal surfaces and reaction pathways. J. Chem. Phys. 2014, 140, 84106. [Google Scholar] [CrossRef] [Green Version]
  51. Li, H.; Shen, Y.Y.; Du, H.N.; Li, J.; Zhang, H.X.; Xu, C.X. Insight into the mechanisms of CO2 reduction to CHO over Zr-doped Cu nanoparticle. Chem. Phys. 2021, 540, 111012. [Google Scholar] [CrossRef]
  52. Chuang, F.C.; Wang, C.Z.; Ho, K.H. Structure of neutral aluminum clusters Aln (2 ≤ n ≤ 23): Genetic algorithm tight-binding calculations. Phys. Rev. B-Condens. Matter Mater. Phys. 2006, 73, 125431. [Google Scholar] [CrossRef]
  53. Mao, H.Y.; Li, B.X.; Ding, W.F.; Zhu, Y.H.; Yang, X.X.; Li, C.Y.; Ye, G.X. Theoretical study on the aggregation of copper clusters on a liquid surface. Materials 2019, 12, 3877. [Google Scholar] [CrossRef]
  54. Holec, D.; Dumitraschkewitz, P.; Fischer, F.D.; Vollath, D. Size-dependent surface energies of Au nanoparticles. arXiv 2014, arXiv:1412.7195. [Google Scholar]
  55. Shin, D.Y.; Won, J.S.; Kwon, J.A.; Kim, M.S.; Lim, D.H. First-principles study of copper nanoclusters for enhanced electrochemical CO2 reduction to CH4. Comput. Theor. Chem. 2017, 1120, 84–90. [Google Scholar] [CrossRef]
  56. Vogt, C.; Monai, M.; Sterk, E.B.; Palle, J.; Melcherts, A.E.M.; Zijlstra, B.; Groeneveld, E.; Berben, P.H.; Boereboom, J.M.; Hensen, E.J.M.; et al. Understanding carbon dioxide activation and carbon–carbon coupling over nickel. Nat. Commun. 2019, 10, 5330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Alvarez, S. A cartography of the van der Waals territories. Dalt. Trans. 2013, 42, 8617–8636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Cordero, B.; Gómez, V.; Platero-Prats, A.E.; Revés, M.; Echeverría, J.; Cremades, E.; Barragán, F.; Alvarez, S. Covalent radii revisited. J. Chem. Soc. Dalt. Trans. 2008, 21, 2832–2838. [Google Scholar] [CrossRef]
Figure 1. (a) The 55-atom Cu-based clusters doped with one metal atom (M) at three different coordination sites: CN = 6, 8 and 12. (b) The 55-atom Cu-based cluster doped with 12 metal atoms located at CN = 6. (c) The 55-atom Cu-based cluster doped with 30 metal atoms at CN = 8. (d) The core(Cu)@shell(M) cluster model.
Figure 1. (a) The 55-atom Cu-based clusters doped with one metal atom (M) at three different coordination sites: CN = 6, 8 and 12. (b) The 55-atom Cu-based cluster doped with 12 metal atoms located at CN = 6. (c) The 55-atom Cu-based cluster doped with 30 metal atoms at CN = 8. (d) The core(Cu)@shell(M) cluster model.
Nanomaterials 13 00087 g001
Figure 2. The segregation energy of the metal dopants Ag, Cd, Pd, Pt, and Zn in Cu54M clusters. Dopant metal located at two different coordination sites: CN6 and CN8.
Figure 2. The segregation energy of the metal dopants Ag, Cd, Pd, Pt, and Zn in Cu54M clusters. Dopant metal located at two different coordination sites: CN6 and CN8.
Nanomaterials 13 00087 g002
Figure 3. The structures and adsorption energies (in eV) of CO2 on the CuM clusters.
Figure 3. The structures and adsorption energies (in eV) of CO2 on the CuM clusters.
Nanomaterials 13 00087 g003
Scheme 1. Reaction pathways to the C1 products CO, HCOOH, CH2O, CH4 and CH3OH.
Scheme 1. Reaction pathways to the C1 products CO, HCOOH, CH2O, CH4 and CH3OH.
Nanomaterials 13 00087 sch001
Figure 4. The Gibbs free energy diagram for the (a) CO2 reduction reaction to CO and the (b) hydrogen evolution reaction on Cu55 cluster and on the (100), (110) and (111) Cu surfaces.
Figure 4. The Gibbs free energy diagram for the (a) CO2 reduction reaction to CO and the (b) hydrogen evolution reaction on Cu55 cluster and on the (100), (110) and (111) Cu surfaces.
Nanomaterials 13 00087 g004
Figure 5. The Gibbs free energy diagram for the CO2 reduction to CO (left), CO2 reduction to HCOOH (centre), and hydrogen evolution reaction (right) on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Figure 5. The Gibbs free energy diagram for the CO2 reduction to CO (left), CO2 reduction to HCOOH (centre), and hydrogen evolution reaction (right) on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Nanomaterials 13 00087 g005
Figure 6. The Gibbs free energy diagram for the eCO2RR pathways to HCHO (left) and CH3OH (right) on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Figure 6. The Gibbs free energy diagram for the eCO2RR pathways to HCHO (left) and CH3OH (right) on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Nanomaterials 13 00087 g006
Figure 7. The limiting potential difference (∆∆GPLS) between CO2RR to CH4 and HER on the bimetallic CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Figure 7. The limiting potential difference (∆∆GPLS) between CO2RR to CH4 and HER on the bimetallic CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Nanomaterials 13 00087 g007
Figure 8. The overpotentials (η) for the electrocatalytic formation of H2, CO, HCOOH, CH2O, CH3OH and CH4 on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Figure 8. The overpotentials (η) for the electrocatalytic formation of H2, CO, HCOOH, CH2O, CH3OH and CH4 on the CuM clusters (M = Ag, Cd, Pd, Pt and Zn): (a) Cu54M with CN6, (b) Cu54M with CN8, (c) Cu43M12, (d) Cu25M30 and (e) core@shell.
Nanomaterials 13 00087 g008
Table 1. The bond lengths (Å), formation energy (eV), HOMO-LUMO gap (ΔH-L, eV), Bader charge difference (∆QM, Coulomb), and surface energy (γ, eV) of the Cu and CuM (M = Ag, Cd, Pd, Pt and Zn) nanoclusters.
Table 1. The bond lengths (Å), formation energy (eV), HOMO-LUMO gap (ΔH-L, eV), Bader charge difference (∆QM, Coulomb), and surface energy (γ, eV) of the Cu and CuM (M = Ag, Cd, Pd, Pt and Zn) nanoclusters.
Bond LengthFormation EnergyΔH−L∆QMγ
Pristine Cu55 Nanocluster
Cu552.51−2.99−0.56
1-atom doping on CN6
Cu54Ag12.69−3.510.0001−0.12−0.56
Cu54Cd12.73−3.480.04120.16−0.54
Cu54Pd12.59−3.530.1013−0.37−0.58
Cu54Pt12.56−3.560.0441−0.64−0.59
Cu54Zn12.54−3.480.00050.13−0.58
1-atom doping on CN8
Cu54Ag12.69−3.520.0488−0.08−0.56
Cu54Cd12.73−3.490.03490.14−0.54
Cu54Pd12.59−3.540.0431−0.30−0.58
Cu54Pt12.56−3.570.0481−0.63−0.59
Cu54Zn12.54−3.490.04140.17−0.58
12-atom doping on CN6
Cu43Ag122.67−3.300.0858−0.12−0.51
Cu43Cd122.75−2.900.01040.14−0.42
Cu43Pd122.59−3.590.0682−0.32−0.64
Cu43Pt122.56−3.960.0824−0.58−0.68
Cu43Zn122.54−2.980.17210.09−0.49
30-atom doping on CN8
Cu-MM-M
Cu25Ag302.652.81−2.930.0305−0.06−0.46
Cu25Cd302.673.01−1.950.00160.08−0.26
Cu25Pd302.592.70−3.490.0568−0.15−0.71
Cu25Pt302.592.69−4.450.0524−0.26−0.82
Cu25Zn302.532.75−2.120.09800.09−0.34
Core@shell
Cu13Ag422.81−2.810.0835−0.14−0.40
Cu13Cd422.95−1.310.0131−0.93−0.15
Cu13Pd422.68−3.400.01680.78−0.72
Cu13Pt422.62−4.720.05270.69−0.86
Cu13Zn422.56−1.530.0173−0.93−0.23
Table 2. The adsorption energies (Eads, eV), bond angles (θOCO, °), charge difference (∆QM, Coulomb) bond elongations (∆lCO, Å) and d-bands center (δd, eV) of CO2 adsorbed on the Cu-M nanoclusters (M = Ag, Cd, Pd, Pt and Zn). The symbols η(Cu, C) and η(M, C) refer to the coordination of the C atom of the adsorbed CO2 molecule to the Cu and M atoms, respectively.
Table 2. The adsorption energies (Eads, eV), bond angles (θOCO, °), charge difference (∆QM, Coulomb) bond elongations (∆lCO, Å) and d-bands center (δd, eV) of CO2 adsorbed on the Cu-M nanoclusters (M = Ag, Cd, Pd, Pt and Zn). The symbols η(Cu, C) and η(M, C) refer to the coordination of the C atom of the adsorbed CO2 molecule to the Cu and M atoms, respectively.
Eadsθ(O-C-O)∆QM∆lCO δd
η(Cu,C)η(M,C)η(Cu,C)η(M,C)η(Cu,C)η(M,C)
Cu55−0.0133.400.05−2.27
1-atom doping (CN6)
Cu54Ag1−0.20−0.200.3400.4600.040.040.00−2.28
Cu54Cd10.07−0.1843.840.2500.610.040.09−2.38
Cu54Pd1−0.20−0.330.71040.120.040.510.17−2.29
Cu54Pt1−0.19−0.560.75044.670.040.570.21−2.29
Cu54Zn1−0.13−0.2047.450.3000.700.040.23−2.36
1-atom doping (CN8)
Cu54Ag1−0.18−0.190.650.420.040.040.00−2.28
Cu54Cd10.14−0.1741.30.410.580.050.15−2.34
Cu54Pd1−0.18−0.220.7245.520.050.630.17−2.28
Cu54Pt1−0.21−0.431.2850.970.630.710.24−2.26
Cu54Zn10.050.2648.852.210.760.860.27−2.34
12-atom doping (CN6)
Cu43Ag12−0.21−0.210.6600.3000.040.040.01−2.67
Cu43Cd12−0.44−0.2348.5045.600.750.680.23−4.03
Cu43Pd12−0.22−0.250.47038.020.040.490.19−2.12
Cu43Pt12−0.07−0.341.26042.380.780.530.29−2.05
Cu43Zn12−0.220.1349.2249.300.750.770.27−3.72
30-atom doping (CN8)
Cu25Ag30−0.22−0.230.4900.6200.040.040.00−3.12
Cu25Cd30−0.09−0.6432.190.2400.650.050.27−5.27
Cu25Pd300.00−0.5235.3042.330.380.480.25−1.49
Cu25Pt30−0.34−0.561.07044.720.020.490.29−1.75
Cu25Zn30−0.50−0.3749.6850.850.750.820.23−4.61
42-atom doping (core@shell)
Cu13Ag42−0.15−0.210.370.340.030.040.00−3.56
Cu13Cd42−0.11−0.200.750.460.040.060.00−7.29
Cu13Pd42−0.13−0.221.962.040.040.040.00−1.58
Cu13Pt42−0.10−0.170.821.090.020.030.00−1.92
Cu13Zn42−0.20−0.180.100.410.020.040.00−6.23
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

Nabi, A.G.; Aman-ur-Rehman; Hussain, A.; Chass, G.A.; Di Tommaso, D. Optimal Icosahedral Copper-Based Bimetallic Clusters for the Selective Electrocatalytic CO2 Conversion to One Carbon Products. Nanomaterials 2023, 13, 87. https://doi.org/10.3390/nano13010087

AMA Style

Nabi AG, Aman-ur-Rehman, Hussain A, Chass GA, Di Tommaso D. Optimal Icosahedral Copper-Based Bimetallic Clusters for the Selective Electrocatalytic CO2 Conversion to One Carbon Products. Nanomaterials. 2023; 13(1):87. https://doi.org/10.3390/nano13010087

Chicago/Turabian Style

Nabi, Azeem Ghulam, Aman-ur-Rehman, Akhtar Hussain, Gregory A. Chass, and Devis Di Tommaso. 2023. "Optimal Icosahedral Copper-Based Bimetallic Clusters for the Selective Electrocatalytic CO2 Conversion to One Carbon Products" Nanomaterials 13, no. 1: 87. https://doi.org/10.3390/nano13010087

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

Article Metrics

Back to TopTop