Next Article in Journal
Electrical and Magnetic Transport Properties of Co2VGa Half-Metallic Heusler Alloy
Previous Article in Journal
Non-Precious Metal Alloy Double Crown-Retained Removable Partial Dentures: A Cross-Sectional In Vivo Investigation
Previous Article in Special Issue
Molecular Dynamics Investigation of Spreading Performance of Physiological Saline on Surface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Adsorption over 3d Transition-Metal Nanoclusters Supported on Pyridinic N3-Doped Graphene: A DFT Investigation

by
Fernando Montejo-Alvaro
1,†,
Jesus A. Martínez-Espinosa
2,†,
Hugo Rojas-Chávez
3,
Diana C. Navarro-Ibarra
1,
Heriberto Cruz-Martínez
1,* and
Dora I. Medina
2,*
1
Tecnológico Nacional de México, Instituto Tecnológico del Valle de Etla, Abasolo S/N, Barrio del Agua Buena, Santiago Suchilquitongo, Oaxaca 68230, Mexico
2
Tecnologico de Monterrey, School of Engineering and Sciences, Atizapán de Zaragoza, Estado de México 52926, Mexico
3
Tecnológico Nacional de México, Instituto Tecnológico de Tláhuac II, Camino Real 625, Col. Jardines del Llano, San Juan Ixtayopan, Alcaldía Tláhuac, Ciudad de México 13550, Mexico
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 18 July 2022 / Revised: 24 August 2022 / Accepted: 30 August 2022 / Published: 4 September 2022
(This article belongs to the Special Issue Modelling Materials and Devices at Atomistic Level)

Abstract

:
CO2 adsorption on bare 3d transition-metal nanoclusters and 3d transition-metal nanoclusters supported on pyridinic N3-doped graphene (PNG) was investigated by employing the density functional theory. First, the interaction of Co13 and Cu13 with PNG was analyzed by spin densities, interaction energies, charge transfers, and HUMO-LUMO gaps. According to the interaction energies, the Co13 nanocluster was adsorbed more efficiently than Cu13 on the PNG. The charge transfer indicated that the Co13 nanocluster donated more charges to the PNG nanoflake than the Cu13 nanocluster. The HUMO-LUMO gap calculations showed that the PNG improved the chemical reactivity of both Co13 and Cu13 nanoclusters. When the CO2 was adsorbed on the bare 3d transition-metal nanoclusters and 3d transition-metal nanoclusters supported on the PNG, it experienced a bond elongation and angle bending in both systems. In addition, the charge transfer from the nanoclusters to the CO2 molecule was observed. This study proved that Co13/PNG and Cu13/PNG composites are adequate candidates for CO2 adsorption and activation.

1. Introduction

Due to environmental changes such as rising global temperatures, sea levels, and melting of the polar ice caps, concerns are accumulating around the presence of CO2 in the atmosphere [1]. Furthermore, population growth and rapid industrialization have escalated the release of CO2 into the atmosphere [2]. In addition, energy requirements for human activities (anthropogenic) have contributed substantially to net CO2 emissions [3]. In fact, anthropogenic CO2 emissions have dramatically increased from 27 Gt in 1970 to 49 Gt in 2010 [4]. As a result, many governments have signed the Kyoto Protocol of the United Nations Framework Convention on Climate Change as a precautionary measure to mitigate climate change [5].
Additionally, several technological processes have been implemented to curb CO2 accumulation by implementing techniques such as the CO2 capturing and sequestration process [6], CO2 capturing and utilization processes [7], and CO2 conversion [8,9]. Yet, the enormous dissociation energy of CO2 resulting from its molecular stability limits the feasibility of the latter process [10]. Moreover, challenges in developing suitable catalysts in terms of efficiency and selectivity still stand unresolved [11]. That notwithstanding, strides in this area continually seek to find more efficient catalysts that diminish process energy requirements and show high selectivity and stability to CO2 reduction reactions.
In this regard, a wide variety of catalysts have been studied: primarily, current studies have been centered on nanoclusters, since they exhibit superior catalytic activity with respect to bulk materials handling, enabled by their specific surface area, electronic, optical, magnetic, and mechanical properties, which are strongly associated with their shape, size, and composition [12]. Thus, 3d transition-metal nanoclusters have gained considerable attention since they have demonstrated remarkable performance in CO2 reduction. For instance, the removal and conversion of CO2 into liquid fuels using amorphous Cu nanoparticles showed an excellent catalytic activity and selectivity [13]. In another study, Li and colleagues studied the CO2 reduction over Ni-based electrocatalysts with sizes ranging from a few atoms to over 100 nm. Their results imply that CO2 reduction and selectivity performance vary according to the size of Ni nanoparticles [14]. Furthermore, another study used Co nanoparticles ranging from 10 to 50 nm prepared for CO2 methanation and ethanol reforming, demonstrating the nanoparticles’ significant activity in the process [15].
Though the catalytic activities of nanoclusters are acclaimed, they tend to agglomerate, resulting in a decline in activity and stability [16,17,18]. Therefore, addressing this issue requires the use of high-surface-area nanomaterials to anchor nanoparticles to promote stability and activation. One of the most eminent nanomaterials in this respect is graphene which exhibits high mechanical, thermal, and electrical properties [19,20,21]. Nonetheless, graphene often necessitates modifications such as doping, functionalization, or the presence of defects on its surface. These adjustments are intended to overcome the low chemical activity of graphene [21,22]. Among the different modifications made to graphene, the pyridinic N3-doped graphene (PNG) has acquired great importance since it substantially modifies the structural and electronic properties of the pristine graphene. Furthermore, studies based on density functional theory (DFT) have proved that the stability and reactivity of metal nanoclusters were enhanced using PNG layers as support [23,24,25].
Moreover, using 3d transition-metal nanoclusters supported on PNG for CO2 reduction has been studied experimentally. For example, Dongare and collaborators investigated electrochemical CO2 reduction using N-doped graphene-supported Cu nanoparticles. The presence of pyridinic, pyrrolic, and graphitic N were confirmed using X-ray photoelectron spectroscopy (XPS). Furthermore, the high PNG content signified good selectivity toward CO2 reduction. In the end, the N-doped graphene/Cu nanoparticles composite enhanced the activity and selectivity in CO2 reduction [26]. Likewise, the incorporation of MnO nanoparticles into an N-doped graphene aerogel for CO2-to-CO electrochemical reduction was studied. According to XPS spectra, it was found that the primary N species was the pyridinic type. Their results suggest that CO2-to-CO electrochemical reduction was enhanced by the synergistic effect of the MnO nanoparticles and the N-doped aerogels [27]. Thus, these studies emphasize the central role of N-doped graphene in CO2 reduction. However, there are no theoretical studies that analyze the effect of the PNG support on the stability and catalytic activity of 3d transition-metal clusters toward CO2 at the molecular level. Therefore, in this work, CO2 adsorption and activation on bare 3d transition-metal nanoclusters and 3d transition-metal supported nanoclusters were investigated using DFT calculations, since their adsorption and activation on surfaces are key steps in the CO2 conversion reaction [28]. To analyze the CO2 interaction on both bare nanoclusters and nanoclusters supported on PNG, the CO2 adsorption energy, CO2 bond elongation, CO2 bending angle, and charge transfer are calculated since these are indicators of effective CO2 dissociation [29,30,31,32].

2. Materials and Methods

To obtain the most stable CO2 adsorption on both the bare nanoclusters and nanoclusters supported on PNG graphene, nine initial structures were considered. All initial structures were optimized using the auxiliary density functional theory (ADFT) method implemented in the deMon2k 4.3.8 software [33]. The ADFT is a reliable and efficient alternative to the conventional DFT approach that allows calculations of large complex systems with less computational effort. The revised Perdew–Burke–Ernzerhof (revPBE) functional was used as the exchange–correlation functional [34]. For the Co, H, C, O, and N atoms, a double-zeta valence plus polarization basis set optimized for generalized gradient approximation (GGA) functionals (DZVP-GGA) was employed [35], and a triple-zeta valence plus polarization basis set optimized for GGA functionals (TZVP-GGA) was used for the Cu atoms [35]. The most stable structures obtained with the ADFT method were reoptimized with the conventional DFT method using the Orca 5.0 software [36]. All DFT calculations were performed using revPBE [34]. The Ahlrichs basis set def2-SVP was used for the C, N, O, and H atoms, and def2-TZVP was used for the Co and Cu atoms [37]. The energy change = 5 × 10−6 Eh, max. gradient = 3 × 10−4 Eh/Bohr, RMS gradient = 1 × 10−4 Eh/Bohr, max. displacement = 4 × 10−3 Bohr, and RMS displacement = 2 × 10−3 Bohr were the convergence criteria used for geometry optimization.
The icosahedral Co13 and Cu13 nanoclusters (Figure 1a) and the graphene model (Figure 1b) used in this work are reported in Figure 1.
The CO2 adsorption energies (Eads) on isolated nanoclusters and nanoclusters supported on PNG were calculated using the basis set superposition error (BSSE) [38]. Moreover, the atom-pairwise (atom-triplewise) dispersion (D3) correction was used for the RevPBE functional using optimized parameters by Grimme et al. [39]. Finally, to analyze the molecular interactions of the Co13 and Cu13 nanoclusters supported on PNG and the CO2 adsorption over bare nanoclusters and nanoclusters supported on PNG, the quantum theory of atoms in molecules implemented in the Multiwfn program was used [40].

3. Results

3.1. Properties of Co13, Cu13, Co13/PNG, and Cu13/PNG

First, comparison of the spin multiplicities of bare nanoclusters and nanoclusters supported on the PNG was made. The Co13 and Cu13 nanoclusters possess spin multiplicity values of 32 and 6, respectively, which agree with previous results reported for the Co13 [41] and Cu13 [42] nanoclusters. The optimized Co13 and Cu13 nanoclusters supported on PNG are illustrated in Figure 2. When the Co13 and Cu13 nanoclusters are supported on PNG, a decrement in spin multiplicity was observed in the supported nanoclusters compared to the bare nanoclusters. The Co13/PNG composite ended with a spin multiplicity value of 27, whereas the Cu13/PNG composite was singlet. The spin multiplicity diminution observed in both systems is ascribed to the stabilizing effect of the PNG. This is also consistent with data reported in literature. For instance, for the Pd- and Pt-based nanoclusters supported on the PNG nanoflake, the latter lead to a magnetic-moment decrement in most of the systems; such behavior was associated with the charge transferred from the nanoclusters to the PNG nanoflake [23,24].
Spin densities, interaction energies (Eint), and charge transfers were calculated to better understand the interaction between the nanoclusters and PNG support. In Figure 3 the spin density of the Co13 nanocluster supported on the PNG is illustrated, since it is an open-shell system, and the spin density of the Co13 nanocluster is also computed for comparison. As can be seen in Figure 3a, for the Co nanocluster, there is a homogeneous spin density distribution over the Co atoms. When the Co nanocluster is supported on the PNG, the spin density is distributed mostly on the Co atoms and little on the N atoms (Figure 3b).
The Eint calculated for the Co13 and Cu13 nanoclusters supported on PNG are −5.69 eV and −4.72 eV, respectively. It should be noted that both values are substantially higher than those reported in previous findings for the Co13 [43,44] and Cu13 [45,46] nanoclusters supported on pristine graphene. Therefore, PNG has a more substantial stabilizing effect over the Co13 and Cu13 nanoclusters compared to pristine graphene. The interaction between the Co13 and Cu13 nanoclusters with PNG was further investigated by Bader charge transfer. The charge transfer between the Co13 and Cu13 nanoclusters and the PNG is reported in Table 1. The results suggest that both nanoclusters yielded charge to the PNG since both ended with a total positive charge. The large electronegativity of N and the electronegativity difference of Co and Cu atoms are the principal driving force for the charge transfer. Furthermore, the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) gap was computed for the bare Co13 and Cu13 nanoclusters and the Co13/PNG and Cu13/PNG composites, Table 1. The Co13 and Cu13 nanoclusters show HOMO-LUMO gaps of 0.52 and 0.67 eV, respectively, which imply that the Co13 can be more reactive than the Cu13 nanocluster as it has been documented that a low HOMO-LUMO gap is associated with high chemical reactivity [45,47,48]. The HOMO-LUMO gaps of the Co13/PNG (0.26 eV) and Cu13/PNG (0.21 eV) were lower than that for the bare nanoclusters. Therefore, it is assumed that the presence of the PNG increased the chemical reactivity of the nanoclusters.

3.2. CO2 Adsorption on Co13 and Cu13 Nanoclusters

Several modes of CO2 adsorption on both the bare nanoclusters and nanoclusters supported on PNG were investigated. First, the CO2 molecule was placed over different facets of the Co13 and Cu13 nanoclusters to identify the most optimal adsorption modes. Due to the high symmetry of the icosahedral Co13 and Cu13 nanoclusters, only nine adsorption modes were considered (Figure 4). Specific parameters were considered to evaluate the CO2 adsorption on the Co13 and Cu13 nanoclusters, such as the Eads, the average bond length, the bending angle of the CO2 molecule, and the charge transfer from the nanoclusters to the CO2 molecule. These parameters were calculated over the most stable structures of CO2 adsorption on nanoclusters and nanoclusters supported on PNG.
The most stable model of CO2 adsorption on the Co13 and Cu13 nanocluster is depicted in Figure 5a and Figure 5b, respectively. The results show that CO2 bends when it is adsorbed on the Co13 and Cu13 nanoclusters. Although the CO2 interaction on the Co13 nanocluster is consistent with the findings of other works [30], the interaction of CO2 and Cu13 differs from the one reported in previous studies, because it was reported that CO2 remained linear once adsorbed and only interacted with one Cu atom [30]. However, this distinct interaction might be attributed to the fact that they used nonicosahedral Cu13 nanoclusters.
To estimate the interaction between the CO2 molecule and the 3d transition-metal nanoclusters, the Eads were calculated. In the case of the CO2 adsorbed on the Co13 nanocluster, the Eads was −0.94 eV. This value is like those reported in the literature [30,49]. The Eads calculated for CO2 on the Cu13 nanocluster was −0.18 eV, which agrees with the data reported in literature [30,49]. In addition, an average bond elongation of the CO2 was observed when it was adsorbed on the Co13 nanocluster. Regarding the bending angle, there was a pronounced curving with a reduction of 23.94% with respect to the isolated CO2 molecule. These parameters are comparable to those reported in previous works [30]. Concerning the average bond length and bending angle of the CO2 molecule adsorbed on the Cu13 nanocluster. As in the CO2 adsorbed on the Co13 nanocluster, an average bond elongation was observed. On the other hand, the bending angle underwent a considerable contraction of 28.34%. Nevertheless, these parameters do not concur with the values reported in earlier findings. In fact, Ocampo-Restrepo did not report this kind of modification in either the average bond length or bending angle of the CO2 molecule [30].

3.3. CO2 Adsorption on Co13 and Cu13 Nanoclusters Supported on PNG

To investigate the CO2 adsorption on Co13/PNG and Cu13/PNG composites, we analyzed various adsorption modes described in the preceding section. Again, the Eads, the average bond length, and the bending angle were determined to evaluate the CO2 adsorption. The most optimal adsorption mode for Co13/PNG and Cu13/PNG composites is depicted in Figure 6a and Figure 6b, respectively. For both composites, the CO2 preferred to be adsorbed in a bent way. Subsequently, the Eads of CO2 on the Co13/PNG and Cu13/PNG composites was calculated. The results showed that the CO2 adsorbs more intensely on the Co13/PNG composite with a value of −0.92 eV, whereas the Eads of CO2 on the Cu13/PNG composite was only −0.33 eV. As in the CO2 adsorbed on bare clusters, an average bond elongation of CO2 was observed when it was adsorbed on nanoclusters supported on PNG (Figure 6). Moreover, the CO2 bending angle underwent a considerable contraction when it was adsorbed on Co13 (24.72%) and Cu13 (22.83%) supported on PNG.

3.4. Bonding Analysis of the CO2 Interaction on Nanoclusters and Nanoclusters Supported on PNG

Table 2 shows the CO2 charge transference when is adsorbed on Co13, Cu13, Co13/PNG, and Cu13/PNG. The total charge of the CO2 molecule resulted in negative values for all systems, which indicates that the CO2 molecule gained a charge after the adsorption. For instance, the CO2 adsorption on the Co13 nanocluster enabled the CO2 to gain a charge of −0.74 e, indicating that the charge is transferred from the nanocluster to the CO2 molecule. It was reported that the transfer plays a significant role in the activation of the CO2 molecule [32,49].
The electron localization function (ELF) was used to analyze the type of bond formed by the CO2 interaction on free nanoclusters and nanoclusters supported on PNG. ELF analysis is widely used to understand the nature of chemical bonds, which allows for differentiating covalent bonds and lone pairs (ELF = 1), as well as ionic bonds, hydrogen bonds, and van der Waals interactions (ELF = 0). Figure 7 shows the ELF plots for the CO2 adsorbed on nanoclusters and nanoclusters supported on PNG. As can be observed, the red region between the C atoms and the metal atom has ELF values of 0.8 and 0.9 for the C-Co (Figure 7a) and C-Cu bonds (Figure 7b), respectively. The high ELF value indicates a high charge transfer from the Cu13 to the CO2 molecules, in accordance with the total charge calculated (see Table 2). For the CO2 absorbed on the nanocluster/PNG, the ELF value is ~0.85 when the CO2 is absorbed on the Co13/PNG composite and ~0.8 when the CO2 is absorbed on the Cu13/PNG composite, suggesting a lower charge transfer from the metal cluster to the CO2 molecule (see Figure 7c,d). For the CO2 adsorption on Co13 and on Co13/PNG, a red region between the C and Co is observed, suggesting a bonding C-Co. Meanwhile, for CO2 on Cu13 a uniform red region is observed between C and a Cu. However, when CO2 is adsorbed on Cu13/PNG a semicircular region appears between C and two Cu atoms (see Figure 7d), suggesting that C is bonded with two Cu atoms.

4. Conclusions

The CO2 adsorption on bare 3d transition-metal nanoclusters and 3d transition-metal nanoclusters supported on the PNG was investigated using the ADFT. First, the interaction between Co13 and Cu13 nanoclusters with PNG was studied by spin densities, interaction energies, charge transfers, and HUMO-LUMO gaps. The results revealed that the PNG enhanced the stability and chemical reactivity of the 3d transition-metal nanoclusters. Subsequently, the CO2 was adsorbed on bare 3d transition-metal nanoclusters, and 3d transition-metal nanoclusters supported PNG. Numerous indicators such as bond elongation, angle bending, and charge transfer were used to characterize the CO2 interaction on these systems. In terms of bond elongation, bare nanoclusters and nanoclusters supported on PNG induced a CO2 bond elongation. In addition, the CO2 molecule experienced a bending angle when it was adsorbed on bare nanoclusters and nanoclusters supported on PNG. Charge transfer analysis revealed that the CO2 gained charge when it was adsorbed on 3d transition-metal nanoclusters and 3d transition-metal nanoclusters supported on PNG nanosheet. Although the nanoclusters and nanoclusters supported on PNG exhibited similar reactivity towards CO2, nanoclusters supported on PNG have the advantage of the support material providing excellent stability to the nanoclusters; therefore, they will not present agglomeration problems.

Author Contributions

Conceptualization, F.M.-A., J.A.M.-E., H.C.-M. and D.I.M.; methodology, F.M.-A., J.A.M.-E., H.R.-C., D.C.N.-I. and D.I.M.; formal analysis, F.M.-A., J.A.M.-E., H.R.-C., D.C.N.-I., H.C.-M. and D.I.M.; investigation, F.M.-A., J.A.M.-E. and H.R.-C.; writing—original draft preparation, F.M.-A., J.A.M.-E., H.R.-C., H.C.-M. and D.I.M.; writing—review and editing, H.C.-M. and D.I.M.; funding acquisition, F.M.-A., H.C.-M. and D.I.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors appreciate the funding sources provided by the Tecnológico Nacional de México (TecNM) through the grant numbers 15455.22-P and 13559.22-P. The article processing charge was funded by Tecnologico de Monterrey.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rockström, J.; Steffen, W.; Noone, K.; Persson, Å.; Chapin, F.S.; Lambin, E.; Lenton, T.M.; Scheffer, M.; Folke, C.; Schellnhuber, H.J.; et al. Planetary boundaries: Exploring the safe operating space for humanity. Ecol. Soc. 2009, 14, 32. [Google Scholar] [CrossRef]
  2. Zou, C.; Zhao, Q.; Zhang, G.; Xiong, B. Energy revolution: From a fossil energy era to a new energy era. Nat. Gas Ind. B 2016, 3, 1–11. [Google Scholar] [CrossRef]
  3. Wang, S.; Li, G.; Fang, C. Urbanization, economic growth, energy consumption, and CO2 emissions: Empirical evidence from countries with different income levels. Renew. Sustain. Energy Rev. 2018, 81, 2144–2159. [Google Scholar] [CrossRef]
  4. Intergovernmental Panel on Climate Change. Climate Change 2014 Mitigation of Climate Change; Intergovernmental Panel on Climate Change: Geneva, Switzerland, 2014. [Google Scholar]
  5. Maamoun, N. The Kyoto protocol: Empirical evidence of a hidden success. J. Environ. Econ. Manag. 2019, 95, 227–256. [Google Scholar] [CrossRef]
  6. Pastero, L.; Curetti, N.; Ortenzi, M.A.; Schiavoni, M.; Destefanis, E.; Pavese, A. CO2 capture and sequestration in stable Ca-oxalate, via Ca-ascorbate promoted green reaction. Sci. Total Environ. 2019, 666, 1232–1244. [Google Scholar] [CrossRef]
  7. Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.; Xie, B.; Toe, C.Y.; Zhu, X.; Wang, J.; et al. Industrial carbon dioxide capture and utilization: State of the art and future challenges. Chem. Soc. Rev. 2020, 49, 8584–8686. [Google Scholar] [CrossRef]
  8. Xiao, X.; Xu, Y.; Lv, X.; Xie, J.; Liu, J.; Yu, C. Electrochemical CO2 reduction on copper nanoparticles-dispersed carbon aerogels. J. Colloid Interface Sci. 2019, 545, 1–7. [Google Scholar] [CrossRef]
  9. Sirijaraensre, J.; Limtrakul, J. Hydrogenation of CO2 to formic acid over a Cu-embedded graphene: A DFT study. Appl. Surf. Sci. 2016, 364, 241–248. [Google Scholar] [CrossRef]
  10. Alper, E.; Orhan, O.Y. CO2 utilization: Developments in conversion processes. Petroleum 2017, 3, 109–126. [Google Scholar] [CrossRef]
  11. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef]
  12. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  13. Duan, Y.-X.; Meng, F.-L.; Liu, K.-H.; Yi, S.-S.; Li, S.-J.; Yan, J.-M.; Jiang, Q. Amorphizing of Cu Nanoparticles toward Highly Efficient and Robust Electrocatalyst for CO2 Reduction to Liquid Fuels with High Faradaic Efficiencies. Adv. Mater. 2018, 30, e1706194. [Google Scholar] [CrossRef] [PubMed]
  14. Li, Z.; He, D.; Yan, X.; Dai, S.; Younan, S.; Ke, Z.; Pan, X.; Xiao, X.; Wu, H.; Gu, J. Size-Dependent Nickel-Based Electrocatalysts for Selective CO2 Reduction. Angew. Chemie 2020, 132, 18731–18736. [Google Scholar] [CrossRef]
  15. Riani, P.; Garbarino, G.; Cavattoni, T.; Canepa, F.; Busca, G. Unsupported cobalt nanoparticles as catalysts: Effect of preparation method on catalytic activity in CO2 methanation and ethanol steam reforming. Int. J. Hydrogen Energy 2019, 44, 27319–27328. [Google Scholar] [CrossRef]
  16. Naito, M.; Yokoyama, T.; Hosokawa, K.; Nogi, K. Nanoparticle Technology Handbook; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 9780444641106. [Google Scholar]
  17. Gosens, I.; Post, J.A.; De La Fonteyne, L.J.J.; Jansen, E.H.J.M.; Geus, J.W.; Cassee, F.R.; De Jong, W.H. Impact of agglomeration state of nano- and submicron sized gold particles on pulmonary inflammation. Part. Fibre Toxicol. 2010, 7, 37. [Google Scholar] [CrossRef]
  18. Zare, Y.; Rhee, K.Y.; Hui, D. Influences of nanoparticles aggregation/agglomeration on the interfacial/interphase and tensile properties of nanocomposites. Compos. Part B Eng. 2017, 122, 41–46. [Google Scholar] [CrossRef]
  19. Ghosh, S.K.; Calizo, I.; Teweldebrhan, D.; Pokatilov, E.P.; Nika, D.L.; Balandin, A.A.; Bao, W.; Miao, F.; Lau, C.N. Extremely high thermal conductivity of graphene: Prospects for thermal management applications in nanoelectronic circuits. Appl. Phys. Lett. 2008, 92, 151911. [Google Scholar] [CrossRef]
  20. Sur, U.K. Graphene: A Rising Star on the Horizon of Materials Science. Int. J. Electrochem. 2012, 2012, 237689. [Google Scholar] [CrossRef]
  21. Cruz-Martínez, H.; Rojas-Chávez, H.; Montejo-Alvaro, F.; Peña-Castañeda, Y.; Matadamas-Ortiz, P.; Medina, D. Recent Developments in Graphene-Based Toxic Gas Sensors: A Theoretical Overview. Sensors 2021, 21, 1992. [Google Scholar] [CrossRef]
  22. Mali, K.S.; Greenwood, J.; Adisoejoso, J.; Phillipson, R.; De Feyter, S. Nanostructuring graphene for controlled and reproducible functionalization. Nanoscale 2015, 7, 1566–1585. [Google Scholar] [CrossRef] [Green Version]
  23. Sánchez-Rodríguez, E.; Vargas-Hernández, C.; Cruz-Martínez, H.; Medina, D. Stability, magnetic, energetic, and reactivity properties of icosahedral M@Pd12 (M = Fe, Co, Ni, and Cu) core-shell nanoparticles supported on pyridinic N3-doped graphene. Solid State Sci. 2020, 112, 106483. [Google Scholar] [CrossRef]
  24. Wang, Q.; Tian, Y.; Chen, G.; Zhao, J. Theoretical insights into the energetics and electronic properties of MPt12 (M = Fe, Co, Ni, Cu, and Pd) nanoparticles supported by N-doped defective graphene. Appl. Surf. Sci. 2017, 397, 199–205. [Google Scholar] [CrossRef]
  25. Montejo-Alvaro, F.; González-Quijano, D.; Valmont-Pineda, J.A.; Rojas-Chávez, H.; Juárez-García, J.M.; Medina, D.I.; Cruz-Martínez, H. CO2 Adsorption on PtCu Sub-Nanoclusters Deposited on Pyridinic N-Doped Graphene: A DFT Investigation. Materials 2021, 14, 7619. [Google Scholar] [CrossRef] [PubMed]
  26. Dongare, S.; Singh, N.; Bhunia, H. Nitrogen-doped graphene supported copper nanoparticles for electrochemical reduction of CO2. J. CO2 Util. 2020, 44, 101382. [Google Scholar] [CrossRef]
  27. Wang, M.; Zhang, B.; Ding, J.; Xu, N.; Bernards, M.T.; He, Y.; Shi, Y. Three-Dimensional Nitrogen-Doped Graphene Aerogel-Supported MnO Nanoparticles as Efficient Electrocatalysts for CO2 Reduction to CO. ACS Sustain. Chem. Eng. 2020, 8, 4983–4994. [Google Scholar] [CrossRef]
  28. Wang, H.; Nie, X.; Guo, X.; Song, C. A computational study of adsorption and activation of CO2 and H2 over Fe(1 0 0) surface. J. CO2 Util. 2016, 15, 107–114. [Google Scholar] [CrossRef]
  29. Dean, J.; Yang, Y.; Austin, N.; Veser, G.; Mpourmpakis, G. Design of Copper-Based Bimetallic Nanoparticles for Carbon Dioxide Adsorption and Activation. ChemSusChem 2018, 11, 1169–1178. [Google Scholar] [CrossRef]
  30. Ocampo-Restrepo, V.K.; Zibordi-Besse, L.; Da Silva, J.L.F. Ab initio investigation of the atomistic descriptors in the activation of small molecules on 3d transition-metal 13-atom clusters: The example of H2, CO, H2O, and CO2. J. Chem. Phys. 2019, 151, 214301. [Google Scholar] [CrossRef]
  31. Pangh, A. Catalytic cleavage of CO2 on bimetallic Ni4M (M = Ni, Mo, Sc, and Y) nanoclusters: A DFT study. Catal. Commun. 2020, 149, 106245. [Google Scholar] [CrossRef]
  32. Austin, N.; Butina, B.; Mpourmpakis, G. CO2 activation on bimetallic CuNi nanoparticles. Prog. Nat. Sci. 2016, 26, 487–492. [Google Scholar] [CrossRef] [Green Version]
  33. Geudtner, G.; Calaminici, P.; Carmona-Espíndola, J.; del Campo, J.M.; Domínguez-Soria, V.D.; Moreno, R.F.; Gamboa, G.U.; Goursot, A.; Köster, A.M.; Reveles, J.U.; et al. deMon2k. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 548–555. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev. Lett. 1998, 80, 890. [Google Scholar] [CrossRef]
  35. Calaminici, P.; Janetzko, F.; Köster, A.M.; Mejia-Olvera, R.; Zuniga-Gutierrez, B. Density functional theory optimized basis sets for gradient corrected functionals: 3d transition metal systems. J. Chem. Phys. 2007, 126, 044108. [Google Scholar] [CrossRef]
  36. Neese, F. Software update: The ORCA program system, version 4.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, 1–6. [Google Scholar] [CrossRef]
  37. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  38. Boys, S.F.; Bernardi, F.D. The Calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  39. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456. [Google Scholar] [CrossRef]
  40. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  41. Wang, Y.; Wang, L.; Ma, S. Role of defects in tuning the adsorption of CO over graphene-supported Co13 cluster. Appl. Surf. Sci. 2019, 481, 1080–1088. [Google Scholar] [CrossRef]
  42. Montejo-Alvaro, F.; Rojas-Chávez, H.; Herrera-Rivera, R.; Mtz-Enriquez, A.; Cruz-Martínez, H.; Medina, D. Investigating the stability of icosahedral Ni13–xCux (x = 1–12) bimetallic nanoclusters supported on defective graphene: Insights from first-principles calculations. Phys. E Low-dimensional Syst. Nanostructures 2019, 118, 113880. [Google Scholar] [CrossRef]
  43. Liu, L.; Su, Y.; Gao, J.; Zhao, J. Electronic and magnetic properties for Co13 clusters deposited on graphene: A first-principles exploration. Phys. E Low-dimensional Syst. Nanostructures 2012, 46, 6–11. [Google Scholar] [CrossRef]
  44. Alonso-Lanza, T.; Ayuela, A.; Aguilera-Granja, F. Chemical Bonding of Transition-Metal Co13Clusters with Graphene. ChemPhysChem 2015, 16, 3700–3710. [Google Scholar] [CrossRef]
  45. Montejo-Alvaro, F.; Oliva, J.; Zarate, A.; Herrera-Trejo, M.; Hdz-García, H.; Mtz-Enriquez, A. Icosahedral transition metal clusters (M13, M = Fe, Ni, and Cu) adsorbed on graphene quantum dots, a DFT study. Phys. E Low-dimensional Syst. Nanostructures 2019, 110, 52–58. [Google Scholar] [CrossRef]
  46. García-Rodríguez, D.; Mendoza-Huizar, L.; Díaz, C. A DFT study of Cu nanoparticles adsorbed on defective graphene. Appl. Surf. Sci. 2017, 412, 146–151. [Google Scholar] [CrossRef]
  47. Ruiz-Morales, Y. HOMO-LUMO gap as an index of molecular size and structure for polycyclic aromatic hydrocarbons (PAHs) and asphaltenes: A theoretical study. I. J. Phys. Chem. A 2002, 106, 11283–11308. [Google Scholar] [CrossRef]
  48. Caglar, A.; Düzenli, D.; Onal, I.; Tezsevin, I.; Sahin, O.; Kivrak, H. A novel experimental and density functional theory study on palladium and nitrogen doped few layer graphene surface towards glucose adsorption and electrooxidation. J. Phys. Chem. Solids 2020, 150, 109684. [Google Scholar] [CrossRef]
  49. Mendes, P.C.D.; Ocampo-Restrepo, V.K.; Da Silva, J.L.F. Ab initio investigation of quantum size effects on the adsorption of CO2, CO, H2O, and H2 on transition-metal particles. Phys. Chem. Chem. Phys. 2020, 22, 8998–9008. [Google Scholar] [CrossRef]
Figure 1. (a) Icosahedral metal nanoclusters with 13 atoms; (b) Pyridinic N3-doped graphene. Green, black, white, and blue spheres represent metal (Co or Cu), C, H, and N atoms, respectively.
Figure 1. (a) Icosahedral metal nanoclusters with 13 atoms; (b) Pyridinic N3-doped graphene. Green, black, white, and blue spheres represent metal (Co or Cu), C, H, and N atoms, respectively.
Materials 15 06136 g001
Figure 2. Top and side views of the most stable interaction of (a) Co13 and (b) Cu13 nanocluster deposited on pyridinic N3-doped graphene. Green, orange, black, white, and blue spheres represent Co, Cu, C, H, and N atoms, respectively.
Figure 2. Top and side views of the most stable interaction of (a) Co13 and (b) Cu13 nanocluster deposited on pyridinic N3-doped graphene. Green, orange, black, white, and blue spheres represent Co, Cu, C, H, and N atoms, respectively.
Materials 15 06136 g002
Figure 3. Spin density distributions of the systems: (a) Co13 nanocluster and (b) Co13/PNG composite.
Figure 3. Spin density distributions of the systems: (a) Co13 nanocluster and (b) Co13/PNG composite.
Materials 15 06136 g003
Figure 4. CO2 adsorption modes on nanoclusters.
Figure 4. CO2 adsorption modes on nanoclusters.
Materials 15 06136 g004
Figure 5. CO2 adsorbed on (a) Co13 and (b) Cu13 nanoclusters. The average bond length (Å) and angle bending (°) of CO2 molecule is depicted. Green, orange, black, and red spheres represent Co, Cu, C, and O atoms, respectively.
Figure 5. CO2 adsorbed on (a) Co13 and (b) Cu13 nanoclusters. The average bond length (Å) and angle bending (°) of CO2 molecule is depicted. Green, orange, black, and red spheres represent Co, Cu, C, and O atoms, respectively.
Materials 15 06136 g005
Figure 6. CO2 adsorption on (a) Co13/PNG and (b) Cu13/PNG. The bond length (Å) and angle bending (°) of CO2 molecule is depicted. Green, orange, black, white, blue, and red spheres represent Co, Cu, C, H, N, and O atoms, respectively.
Figure 6. CO2 adsorption on (a) Co13/PNG and (b) Cu13/PNG. The bond length (Å) and angle bending (°) of CO2 molecule is depicted. Green, orange, black, white, blue, and red spheres represent Co, Cu, C, H, N, and O atoms, respectively.
Materials 15 06136 g006
Figure 7. Electron localization function contours for: (a) CO2 adsorption on Co13, (b) CO2 adsorption on Cu13, (c) CO2 adsorption on CO13/PNG, and (d) CO2 adsorption on Cu13/PNG.
Figure 7. Electron localization function contours for: (a) CO2 adsorption on Co13, (b) CO2 adsorption on Cu13, (c) CO2 adsorption on CO13/PNG, and (d) CO2 adsorption on Cu13/PNG.
Materials 15 06136 g007
Table 1. Charge transfers and HOMO-LOMO gaps for Co13, Cu13, Co13/PNG, and Cu13/PNG. The positive values indicate charge transfers from the nanoclusters to the PNG structure.
Table 1. Charge transfers and HOMO-LOMO gaps for Co13, Cu13, Co13/PNG, and Cu13/PNG. The positive values indicate charge transfers from the nanoclusters to the PNG structure.
SystemsHOMO-LOMO Gap (eV)Charge Transfer (e)
Co130.52-
Cu130.67-
Co13/PNG0.26+1.40
Cu13/PNG0.21+0.78
Table 2. Charge transfer from Co13, Cu13, Co13/PNG, and Cu13/PNG to CO2 molecule. The negative values indicate charge transfers from the nanoclusters and nanoclusters/PNG to the CO2 molecule.
Table 2. Charge transfer from Co13, Cu13, Co13/PNG, and Cu13/PNG to CO2 molecule. The negative values indicate charge transfers from the nanoclusters and nanoclusters/PNG to the CO2 molecule.
SystemsCharge Transfer toward CO2 (e)
Co13−0.74
Cu13−0.78
Co13/PNG−0.75
Cu13/PNG−0.67
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Montejo-Alvaro, F.; Martínez-Espinosa, J.A.; Rojas-Chávez, H.; Navarro-Ibarra, D.C.; Cruz-Martínez, H.; Medina, D.I. CO2 Adsorption over 3d Transition-Metal Nanoclusters Supported on Pyridinic N3-Doped Graphene: A DFT Investigation. Materials 2022, 15, 6136. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15176136

AMA Style

Montejo-Alvaro F, Martínez-Espinosa JA, Rojas-Chávez H, Navarro-Ibarra DC, Cruz-Martínez H, Medina DI. CO2 Adsorption over 3d Transition-Metal Nanoclusters Supported on Pyridinic N3-Doped Graphene: A DFT Investigation. Materials. 2022; 15(17):6136. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15176136

Chicago/Turabian Style

Montejo-Alvaro, Fernando, Jesus A. Martínez-Espinosa, Hugo Rojas-Chávez, Diana C. Navarro-Ibarra, Heriberto Cruz-Martínez, and Dora I. Medina. 2022. "CO2 Adsorption over 3d Transition-Metal Nanoclusters Supported on Pyridinic N3-Doped Graphene: A DFT Investigation" Materials 15, no. 17: 6136. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15176136

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