Next Article in Journal
Polyfluorene-Based Multicolor Fluorescent Nanoparticles Activated by Temperature for Bioimaging and Drug Delivery
Previous Article in Journal
Computational Modeling of Tensile Stress Effects on the Structure and Stability of Prototypical Covalent and Layered Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability of Coinage Metals Interacting with C60

by
Navaratnarajah Kuganathan
1,2,*,
Ratnasothy Srikaran
3 and
Alexander Chroneos
1,2
1
Department of Materials, Imperial College London, London SW7 2AZ, UK
2
Faculty of Engineering, Environment and Computing, Coventry University, Priory Street, Coventry CV1 5FB, UK
3
Department of Chemistry, University of Jaffna, Sir. Pon Ramanathan Road, Thirunelvely, Jaffna 40000, Srilanka
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(10), 1484; https://doi.org/10.3390/nano9101484
Submission received: 27 September 2019 / Revised: 16 October 2019 / Accepted: 17 October 2019 / Published: 18 October 2019

Abstract

:
Buckminsterfullerene (C60) has been advocated as a perfect candidate material for the encapsulation and adsorption of a variety of metals and the resultant metallofullerenes have been considered for the use in different scientific, technological and medical areas. Using spin-polarized density functional theory together with dispersion correction, we examine the stability and electronic structures of endohedral and exohedral complexes formed between coinage metals (Cu, Ag and Au) and both non-defective and defective C60. Encapsulation is exoergic in both forms of C60 and their encapsulation energies are almost the same. Exohedral adsorption of all three metals is stronger than that of endohedral encapsulation in the non-defective C60. Structures and the stability of atoms interacting with an outer surface of a defective C60 are also discussed. As the atoms are stable both inside and outside the C60, the resultant complexes can be of interest in different scientific and medical fields. Furthermore, all complexes exhibit magnetic moments, inferring that they can be used as spintronic materials.

Graphical Abstract

1. Introduction

Buckminsterfullerene (C60) is one of the allotropes of carbon exhibiting nanosized molecular structure with potential applications in chemistry, catalysis, material science, biology and medicine [1,2,3,4,5,6,7,8]. Encapsulation (atom located inside) and adsorption (atom located outside) of metal atoms have received great experimental and theoretical attention recently in order to optimise the application of C60, mainly in catalytic processes, and electronic and metal storage devices [1,2,3,4,9,10,11,12,13,14,15].
A variety of metal atoms have been encapsulated within C60 experimentally for different applications [1,2,3,4,5,6]. The preparation of endohedral fullerenes can be mainly carried out by inserting metals, either during the preparation of fullerenes in the arc-vaporization technique [16] or after the preparation of fullerenes through five- or six-membered rings [17]. The latter method requires a high kinetic energy barrier as guest atoms need to travel through the five- or six-membered rings. There are other experimental methods available in the literature for the encapsulation [18,19,20]. A laser vaporization technique was used to encapsulate La by Klinger et al. [12] and the electronic behavior of the resultant compound was investigated using tunneling conductivity measurements. A recoil implantation technique was applied to encapsulate radioactive isotopes including 86Zr and 168Hf [17]. The resultant radioactive complexes were suggested for future applications in medical science as direct contact of toxic guest atoms with biological systems in the body can be avoided. Electronic structure calculations based on density functional theory (DFT) on several endohedral fullerenes have been reported in the literature [21,22,23,24]. Recently, we studied the thermodynamical stability and electronic structures of volatile fission atoms (Xe, Kr, Br, I, Te, Rb and Cs) encapsulated within C60 using DFT simulation to predict the efficacy of C60 as a filter material in the spent fuel reprocessing [25].
Although there are many experimental studies available on endohedral fullerenes, only few experimental investigations have been reported on exohedral fullerenes [26,27,28]. Nevertheless, numerous theoretical simulations have been performed to study interactions of metal atoms on the surface of C60 [29,30,31,32]. Exohedral adsorption of a single atom and multiple atoms were studied by Sankar De et al. [29] using ab initio simulation in the absence of dispersion correction. In a very recent DFT simulation, we have shown the importance of dispersion correction for a Cd atom interacting with a C60 surface [33].
Encapsulation of coinage metals by C60 is of great interest as the resultant endohedral compounds can provide useful information about electronic structures of the encapsulated atoms required for the development of electronic materials. Huang et al. [34] synthesized Cu@C60 and its magnetic structure exhibits ring-current-induced magnetism. Experimental formation of Ag@C60 was reported by Narwade et al. [35] and its efficient electrocatalytic activity was tested in the fuel cell reaction. The growth of Au nanoparticles embedded in the C60 cage was experimentally characterised using a thermal co-evaporation technique by Singhal et al. [36]. Although there is an experimental interest in the literature, there are only few theoretical simulations of coinage metals interacting with C60.
In the present study, we used DFT with dispersion correction to examine the stability of coinage metal atoms interacting (Cu, Ag and Au) both inside and outside non-defective and defective C60 molecules. The current simulation method enables the examination of the electronic structures, charge transfer and magnetic moment of the composites.

2. Computational Methods

All calculations were based on the DFT. The VASP code [37,38] which solves the Kohn–Sham (KS) equations using plane wave basis sets was used. The exchange correlation term was modelled using the generalized gradient approximation (GGA) with the Perdew, Burke and Ernzerhof (PBE) function [39]. A plane wave basis set with a cut-off value of 500 eV was used in all cases. Energy minimisation was performed using the conjugate gradient algorithm [40] with a force tolerance of 0.001 eV/Å. A single k-point (Γ) point was used in all calculations to represent the Brillouin zone due to the large supercell. Coinage bulk structures were optimised using a 16 × 16 × 16 Monkhorst k-point mesh [41] which yielded 64 k points. Cubic supercells with 20 Å were used in all calculations to ensure that the two adjacent images do not interact with each other. Encapsulation and adsorption energies were calculated using the following equation:
Eenc / ads = E   ( M C 60 ) E   ( C 60 ) E   ( M ) ,
where E(C60) is the total energy of non-defective or defective C60 molecule, E(M–C60) is the total energy of a metal atom interacting with a non-defective or defective C60 and E (M) is the total energy of an isolated metal atom. Here, dispersive interaction was included by using the pair-wise force field as implemented by Grimme et al. [42] in the VASP package.

3. Results

3.1. Initial Configurations

Six different positions were considered for the interaction of coinage metals with C60. At the endohedral position (endo), an atom occupies the center of a C60 molecule (refer to Figure 1). There are five different positions available on the outer surface of the C60 molecule for the adsorption of atoms, as shown in Figure 1. Positions “hex” and “pent” exhibit that the atom is just above the center of the hexagonal and pentagonal rings of the C60, respectively. In the 66 and 65 configurations, the atom sits on the C−C bond, connecting two adjacent hexagonal rings and connecting one hexagonal and its nearest neighbor pentagonal ring, respectively. In the configuration “C”, the metal atom is adsorbed by a C atom in the C60 cage.

3.2. Validation of the Pseudopotentials and Basis Sets

In order to validate the choice of pseudopotentials and basis sets used in this study for C, Cu, Ag and Au, geometry optimisations were performed on a C60 molecule and bulk structures of coinage metals. Calculated geometrical parameters and electronic structures were then compared with corresponding experimental values and the values reported in other theoretical studies.

3.2.1. Calculation on Band Gap of a C60 Molecule

Geometry optimisation of a C60 molecule revealed bond length values of 1.45 Å and 1.40 Å for C−C and C=C, respectively. These values are in good agreement with the respective values of 1.43 Å and 1.39 Å reported experimentally [43]. The band gap was calculated by plotting the density of states (DOSs) and measuring the distance between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), as shown in the Figure 2. The estimated band gap value of 1.55 eV agrees well with the theoretical value of 1.64 eV [44]. Figure 2 shows the optimised structure of a C60 molecule together with its total DOS plot, HOMO and LUMO.

3.2.2. Calculations of Energy Minimisation on Bulk Cu, Ag and Au

Energy minimisation calculations were performed on bulk Cu, Ag and Au (fcc structure) structures to obtain equilibrium lattice constants and cohesive energies. Cohesive energy was calculated by considering the energy difference between an isolated gas-phase atom and an atom in the bulk. The calculated equilibrium lattice constants and cohesive energies (refer to Table 1) are in good agreement with the experimental and calculated values. Zero magnetic moments were calculated for all three bulk structures. DOS plots for the bulk structures of Cu, Ag and Au (refer to Figure 3) confirm the non-magnetic nature of bulk coinage metals. Considerable progress has recently been made on accurate theoretical determination of electronic band structure of solids. In order to describe the band structure of bulk copper, self-energy of d electrons was included in first-principles GW calculations by Marini et al. [45] and good agreement between experiment and theory was observed. In another theoretical study by Goraus et al. [46], on-site Coulomb interaction (U) was included for Ag 4d states in CeAgGa and the calculated density of states was in satisfactory agreement with X-ray photoemission spectra.

3.3. Encapsulation of Coinage Metals within C60

First, the stability of a single coinage metal occupying the center of the cage was considered. Relaxed structures revealed that the encapsulated atoms (Cu, Ag and Au) are still in the center of the cage without altering their atomic positions. Encapsulation energy was calculated to determine the thermodynamical stability of the atoms when they are inside the C60. The equation showing the methodology of calculating encapsulation energy was reported in the methodology section. Calculations were carried out with and without dispersion. Encapsulation energies calculated using dispersion are exoergic, meaning that they are stable inside the C60 (refer to Table 2). Endoergic encapsulation is observed in calculations without dispersion, inferring the importance of the dispersion. Enhancement in the encapsulation is due to the additional attractive term introduced by the dispersion. The rest of the calculations in this study were only performed with dispersion.
Both Cu and Au have similar encapsulation energies. This can be due to the identical empirical atomic radii of Cu and Au (1.35 Å) although their calculated values are 1.45 Å and 1.74 Å, respectively. The calculated atomic radiis of Cu, Ag and Au are in ascending order. However, the encapsulation energy does not follow any trend with it. Bader charge analysis [49] shows that there is a small amount of charge transferred from metal atoms to the C60 cage. The outer electronic configuration of all three metal atoms is d10s1, meaning that the magnetic moment is one as there is an un-paired electron in the s-shell. The net magnetic moment of a C60 molecule is calculated to be zero. The magnetic moment of the resultant complex is ~1. This indicates that the electronic configuration of the metal is unaltered. This is further supported by the very small positive Bader charge on the metal atoms.
Figure 4 shows the calculated total and atomic DOS plots and partial charge density distribution around the encapsulated atoms. Encapsulation introduces extra peaks associated with s and d orbitals. In the case of Cu, there is a slight overlap between s and d states near the Fermi energy. The d states are further away (~3 eV) from the Fermi energy level while the s states are closer to it. Encapsulation of Au introduces the s states near the Fermi level and the d states are 2 eV away from it.

3.4. Adsorption of Coinage Metals on the Surface of C60

Here, we considered the adsorption of metal atoms by the outer surface of C60. As discussed above, five different initial configurations were considered. Table 3 reports the final configurations and relative adsorption energies (refer to the methodology section for the equation that was used to calculate the adsorption energy) with respect to the most stable configuration.
The configuration “C” was found to be the most stable adsorption site and this is in agreement with the theoretical study reported by Shankar et al. [29]. Figure 5 shows the most stable geometries and bond distances formed between metal atoms and the C60. Calculated adsorption energies are negative, meaning that all three metals are stable on the surface of the C60. Copper forms a shorter Cu–C bond distance of 1.957 Å than the Ag–C60 and Au–C60 bond distances, which reflects in the adsorption energy and the Bader charge (refer to Table 4). The C–Ag bond distance is 2.242 Å which is slightly longer than the C–Cu and C–Au bond distances. The lower adsorption energy can be attributed to the longer C–Ag bond distance. In the case of Au, the adsorption energy is 0.09 eV less than that calculated for Cu. This is evidenced by the intermediate Au–C bond distance of 2.117 Å. Magnetic moments are not altered significantly. However, Bader charge and magnetic moment values are slightly higher and lower than that found in the encapsulated complexes, respectively, confirming the adsorption nature of atoms is stronger than encapsulation.
Figure 6 shows the total DOSs, atomic DOSs and charge density distribution plots showing the interaction of atoms with C60. Additional states arising from s and d orbitals are introduced between the top of the valence band and bottom of the conduction band without altering the band gap significantly.

3.5. Defective C60 Structure

Next, we considered a defective C60 structure to examine the encapsulation or adsorption capability of coinage metal atoms. A defect was introduced in C60 by removing a C atom. The relaxed structure, and DOS and charge density plots of defective C60 are shown in Figure 7. In previous experimental and theoretical studies, non-defective or defective single-walled nanotubes and C60 have been considered for the reaction with transition metals, molecules, one-dimensional crystals and metal clusters [52,53,54,55,56,57].
The defective C60 structure forms an eight-membered ring with distorted C–C bond distances. The C60 molecule has now lost its symmetry and the outer surface of the optimised structure is expected to be more reactive with metal atoms. Furthermore, encapsulation via the eight-membered ring with larger open space can be easier than either a six- or a five-membered ring.

3.6. Encapsulation of Metal Atoms within a Defective C60 Molecule

In the optimised structures, the positions of the atoms are slightly deviated from the center of the cage and towards the defective part of the C60 molecule (refer to Figure 8). This is because of the non-symmetrical nature of the defective C60 molecule. The calculations show that the encapsulation energies are exoergic, meaning that they are more stable in the cage of defective C60 molecule than in isolated atoms (refer to Table 5). Encapsulation energies are approximately −0.50 eV in all cases, indicating that interactions between metal atoms and the defective C60 molecule are almost the same. This is further supported by the similar C‒M bond distances in the relaxed structures (refer to Figure 8). Bader charge analysis shows that metal atoms transfer a small amount of charge to C60. The magnetic moment of the defective C60 molecule is zero. In all cases, magnetic moments of complexes are one, meaning that electronic configurations of coinage metal atoms are almost unaltered.

3.7. Adsorption of Metal Atoms on the Surface of Defective C60 Molecule

Finally, interactions between atoms and defective surface of C60 were considered. There is a strong interaction between defect and metal atoms (refer to Figure 9). Metal atoms occupy the vacant side forming strong bonds with adjacent carbon atoms. This is evidenced by the bond distances, the amount of charge transferred (refer to Table 6), charge density plots and the reduction in the magnetic moments.

4. Conclusions

In this study, DFT simulations, together with dispersion correction, were performed to examine the encapsulation and adsorption capability of both non-defective and defective C60 molecules. Calculations show that the non-defective C60 can trap all three metals both endohedrally and exohedrally. Significant enhancement is observed for the exohedral adsorption compared to for the endohedral encapsulation in the non-defective C60. Furthermore, the defective C60 was examined for trapping both endohedrally and exohedrally. Encapsulation and adsorption energies are exoergic, suggesting that the defective C60 is also a candidate host material for trapping coinage metals. Both non-defective and defective C60 structures can be ideal host materials in scientific and medical fields. As resultant M–C60 complexes are magnetic, they can play an important role in spintronic devices.

Author Contributions

Computation, N.K.; writing of original draft preparation, N.K.; writing of review and editing, R.S. and A.C.

Funding

This research received no external funding.

Acknowledgments

The High Performance Computing (HPC) Center at Imperial College London is acknowledged for providing computational facilities and technical support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kroto, H.W.; Allaf, A.W.; Balm, S.P. C60: Buckminsterfullerene. Chem. Rev. 1991, 91, 1213–1235. [Google Scholar] [CrossRef]
  2. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral Fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  3. Acquah, S.; Penkova, A.V.; Markelov, D.; Semisalova, A.S.; Leonhardt, B.E.; Magi, J.M. Review—The Beautiful Molecule: 30 Years of C60 and Its Derivatives. ECS J. Solid State Sci. Technol. 2017, 6, M3155–M3162. [Google Scholar]
  4. Bühl, M.; Hirsch, A. Spherical Aromaticity of Fullerenes. Chem. Rev. 2001, 101, 1153–1184. [Google Scholar] [CrossRef] [PubMed]
  5. Dresselhaus, M.S.; Dresslhaus, G. Fullerenes and Fullerene Derived Solids as Electronic Materials. Annu. Rev. Mater. Sci. 1995, 25, 487–523. [Google Scholar] [CrossRef]
  6. Bakry, R.; Vallant, R.M.; Najam-ul-Haq, M.; Rainer, M.; Szabo, Z.; Huck, C.W.; Bonn, G.K. Medicinal applications of fullerenes. Int. J. Nanomed. 2007, 2, 639–649. [Google Scholar]
  7. Goodarzi, S.; Da Ros, T.; Conde, J.; Sefat, F.; Mozafari, M. Fullerene: Biomedical engineers get to revisit an old friend. Mater. Today 2017, 20, 460–480. [Google Scholar] [CrossRef]
  8. Kroto, H.W. C60: Buckminsterfullerene, The Celestial Sphere that Fell to Earth. Angew. Chem. Int. Ed. Engl. 1992, 31, 111–129. [Google Scholar] [CrossRef]
  9. He, T.; Gao, G.; Kou, L.; Will, G.; Du, A. Endohedral metallofullerenes (M@C60) as efficient catalysts for highly active hydrogen evolution reaction. J. Catal. 2017, 354, 231–235. [Google Scholar] [CrossRef]
  10. Mahdy, A.M.E. DFT study of hydrogen storage in Pd-decorated C60 fullerene. Mol. Phys. 2015, 113, 3531–3544. [Google Scholar] [CrossRef]
  11. Tawfik, S.A.; Cui, X.Y.; Ringer, S.P.; Stampfl, C. Endohedral metallofullerenes, M@C60 (M = Ca, Na, Sr): Selective adsorption and sensing of open-shell NOx gases. Phys. Chem. Chem. Phys. 2016, 18, 21315–21321. [Google Scholar] [CrossRef] [PubMed]
  12. Klingeler, R.; Kann, G.; Wirth, I.; Eisebitt, S.; Bechthold, P.S.; Neeb, M.; Eberhardt, W. La@C60: A metallic endohedral fullerene. J. Chem. Phys. 2001, 115, 7215–7218. [Google Scholar] [CrossRef]
  13. Cho, S.C.; Kaneko, T.; Ishida, H.; Hatakeyama, R. Nitrogen-atom endohedral fullerene synthesis with high efficiency by controlling plasma-ion irradiation energy and C60 internal energy. J. Appl. Phys. 2015, 117, 123301. [Google Scholar] [CrossRef]
  14. Ohtsuki, T.; Masumoto, K.; Ohno, K.; Maruyma, Y.; Kawazoe, Y.; Sueki, K.; Kikuchi, K. Insertion of Be Atoms in C60 Fullerene Cages: Be@C60. Phys. Rev. Lett. 1996, 77, 3522–3524. [Google Scholar] [CrossRef]
  15. Kuganathan, N.; Green, J.C.; Himmel, H.-J. Dinitrogen fixation and activation by Ti and Zr atoms, clusters and complexes. New J. Chem. 2006, 30, 1253–1262. [Google Scholar] [CrossRef]
  16. Kyesmen, P.I.; Onoja, A.; Amah, A.N. Fullerenes synthesis by combined resistive heating and arc discharge techniques. Springerplus 2016, 5, 1323. [Google Scholar] [CrossRef]
  17. Saha, S.K.; Chowdhury, D.P.; Das, S.K.; Guin, R. Encapsulation of radioactive isotopes into C60 fullerene cage by recoil implantation technique. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2006, 243, 277–281. [Google Scholar] [CrossRef]
  18. Wan, Z.; Christian, J.F.; Anderson, S.L. Collision of Li+ and Na+ with C60: Insertion, fragmentation, and thermionic emission. Phys. Rev. Lett. 1992, 69, 1352–1355. [Google Scholar] [CrossRef]
  19. Saunders, M.; Cross, R.J.; Jiménez-Vázquez, H.A.; Shimshi, R.; Khong, A. Noble gas atoms inside fullerenes. Science 1996, 271, 1693–1697. [Google Scholar] [CrossRef]
  20. Braun, T.; Rausch, H. Endohedral incorporation of argon atoms into C60 by neutron irradiation. Chem. Phys. Lett. 1995, 237, 443–447. [Google Scholar] [CrossRef]
  21. Brocławik, E.; Eilmes, A. Density functional study of endohedral complexes M@C60 (M=Li, Na, K, Be, Mg, Ca, La, B, Al): Electronic properties, ionization potentials, and electron affinities. J. Chem. Phys. 1998, 108, 3498–3503. [Google Scholar] [CrossRef]
  22. Srivastava, A.K.; Pandey, S.K.; Misra, N. Encapsulation of lawrencium into C60 fullerene: Lr@C60 versus Li@C60. Mater. Chem. Phys. 2016, 177, 437–441. [Google Scholar] [CrossRef]
  23. Wang, Y.; Tománek, D.; Ruoff, R.S. Stability of M@C60 endohedral complexes. Chem. Phys. Lett. 1993, 208, 79–85. [Google Scholar] [CrossRef]
  24. Lu, J.; Zhang, X.; Zhao, X. Electronic structures of endohedral N@C60, O@C60 and F@C60. Chem. Phys. Lett. 1999, 312, 85–90. [Google Scholar] [CrossRef]
  25. Kuganathan, N.; Arya, A.K.; Rushton, M.J.D.; Grimes, R.W. Trapping of volatile fission products by C60. Carbon 2018, 132, 477–485. [Google Scholar] [CrossRef]
  26. Breton, J.; Gonzalez-Platas, J.; Girardet, C. Endohedral and exohedral adsorption in C60: An analytical model. J. Chem. Phys. 1993, 99, 4036–4040. [Google Scholar] [CrossRef]
  27. Yekymov, E.; Bounioux, C.; Itzhak-Cohen, R.; Zeiri, L.; Shahnazaryan, E.; Katz, E.A.; Yerushalmi-Rozen, R. All carbon non-covalent exohedral hybrids: C60 aggregates on nanotube networks. J. Energy Chem. 2018, 27, 957–961. [Google Scholar] [CrossRef]
  28. Nakamura, Y.; Kato, S.-I. Exohedral functionalization of fullerenes and supramolecular chemistry. Chem. Rec. 2011, 11, 77–94. [Google Scholar] [CrossRef]
  29. Sankar De, D.; Flores-Livas, J.A.; Saha, S.; Genovese, L.; Goedecker, S. Stable structures of exohedrally decorated C60-fullerenes. Carbon 2018, 129, 847–853. [Google Scholar] [CrossRef]
  30. Özdamar, B.; Boero, M.; Massobrio, C.; Felder-Flesch, D.; Roux, S.L. Exohedral M–C60 and M2–C60 (M = Pt, Pd) systems as tunable-gap building blocks for nanoarchitecture and nanocatalysis. J. Chem. Phys. 2015, 143, 114308. [Google Scholar] [CrossRef]
  31. Kobayashi, T.; Yokoyama, K. Theoretical study of the adsorption of Cs, Cs+, I, I−, and CsI on C60 fullerene. J. Nucl. Sci. Technol. 2016, 53, 1489–1493. [Google Scholar] [CrossRef]
  32. Suzumura, J.-i.; Hosoya, N.; Nagao, S.; Mitsui, M.; Nakajima, A. Electronic structures of exohedral lanthanide-C60 clusters. J. Chem. Phys. 2004, 121, 2649–2654. [Google Scholar] [CrossRef] [PubMed]
  33. Kuganathan, N.; Selvanantharajah, N.; Iyngaran, P.; Abiman, P.; Chroneos, A. Cadmium trapping by C60 and B-, Si-, and N-doped C60. J. Appl. Phys. 2019, 125, 054302. [Google Scholar] [CrossRef]
  34. Huang, H.; Ata, M.; Yoshimoto, Y. Cu@C60 formation in rf-plasma and ring-current induced magnetism of C60. Chem. Commun. 2004, 10, 1206–1207. [Google Scholar] [CrossRef]
  35. Narwade, S.S.; Mulik, B.B.; Mali, S.M.; Sathe, B.R. Silver nanoparticles sensitized C60(Ag@C60) as efficient electrocatalysts for hydrazine oxidation: Implication for hydrogen generation reaction. Appl. Surf. Sci. 2017, 396, 939–944. [Google Scholar] [CrossRef]
  36. Singhal, R.; Sharma, P.; Vishnoi, R.; Avasthi, D.K. Synthesis and characterizations of Au-C60 nanocomposite. J. Alloys Compd. 2017, 696, 9–15. [Google Scholar] [CrossRef]
  37. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  38. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  39. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  40. Press, W.H.; Teukolsky, S.A.; Vetterling, W.T.; Flannery, B.P. Numerical Recipes in C: The Art of Scientific Computing, 2nd ed.; Cambridge University Press: Cambridge, UK, 1992; p. 994. [Google Scholar]
  41. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  42. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  43. Hawkins, J.M. Osmylation of C60: Proof and characterization of the soccer-ball framework. Acc. Chem. Res. 1992, 25, 150–156. [Google Scholar] [CrossRef]
  44. Goclon, J.; Winkler, K.; Margraf, J.T. Theoretical investigation of interactions between palladium and fullerene in polymer. RSC Adv. 2017, 7, 2202–2210. [Google Scholar] [CrossRef] [Green Version]
  45. Marini, A.; Onida, G.; Del Sole, R. Quasiparticle Electronic Structure of Copper in the GW Approximation. Phys. Rev. Lett. 2001, 88, 016403. [Google Scholar] [CrossRef] [PubMed]
  46. Goraus, J.; Ślebarski, A.; Fijałkowski, M. Electronic structure of CeAgGa based on the experimental and model data. Phys. Status Solidi (B) 2011, 248, 2857–2863. [Google Scholar] [CrossRef]
  47. Haas, P.; Tran, F.; Blaha, P. Calculation of the lattice constant of solids with semilocal functionals. Phys. Rev. B 2009, 79, 085104. [Google Scholar] [CrossRef] [Green Version]
  48. Janthon, P.; Luo, S.; Kozlov, S.M.; Viñes, F.; Limtrakul, J.; Truhlar, D.G.; Illas, F. Bulk Properties of Transition Metals: A Challenge for the Design of Universal Density Functionals. J. Chem. Theory Comput. 2014, 10, 3832–3839. [Google Scholar] [CrossRef] [Green Version]
  49. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  50. Slater, J.C. Atomic Radii in Crystals. J. Chem. Phys. 1964, 41, 3199–3204. [Google Scholar] [CrossRef]
  51. Clementi, E.; Raimondi, D.L.; Reinhardt, W.P. Atomic Screening Constants from SCF Functions. II. Atoms with 37 to 86 Electrons. J. Chem. Phys. 1967, 47, 1300–1307. [Google Scholar] [CrossRef]
  52. Zoberbier, T.; Chamberlain, T.W.; Biskupek, J.; Kuganathan, N.; Eyhusen, S.; Bichoutskaia, E.; Kaiser, U.; Khlobystov, A.N. Interactions and Reactions of Transition Metal Clusters with the Interior of Single-Walled Carbon Nanotubes Imaged at the Atomic Scale. J. Am. Chem. Soc. 2012, 134, 3073–3079. [Google Scholar] [CrossRef] [Green Version]
  53. Tascón, J.M.D.; Bottani, E.J. Nitrogen Physisorption on Defective C60. J. Phys. Chem. B 2002, 106, 9522–9527. [Google Scholar] [CrossRef]
  54. Bai, H.; Ji, W.; Liu, X.; Wang, L.; Yuan, N.; Ji, Y. Doping the Buckminsterfullerene by Substitution: Density Functional Theory Studies of C59X (X = B, N, Al, Si, P, Ga, Ge, and As). J. Chem. 2013, 2013, 9. [Google Scholar] [CrossRef]
  55. Kuganathan, N.; Green, J.C. Mercury telluride crystals encapsulated within single walled carbon nanotubes: A density functional study. Int. J. Quantum Chem. 2008, 108, 797–807. [Google Scholar] [CrossRef]
  56. Kuganathan, N.; Green, J.C. Crystal structure of low-dimensional Cu(i) iodide: DFT prediction of cuprophilic interactions. Chem. Commun. 2008, 21, 2432–2434. [Google Scholar] [CrossRef] [PubMed]
  57. Bichoutskaia, E.; Liu, Z.; Kuganathan, N.; Faulques, E.; Suenaga, K.; Shannon, I.J.; Sloan, J. High-precision imaging of an encapsulated Lindqvist ion and correlation of its structure and symmetry with quantum chemical calculations. Nanoscale 2012, 4, 1190–1199. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Initial configurations of a single metal atom interacting with a C60 molecule.
Figure 1. Initial configurations of a single metal atom interacting with a C60 molecule.
Nanomaterials 09 01484 g001
Figure 2. Relaxed structure (a), total density of state (DOS) (b), highest occupied molecular orbital (HOMO) (c), and lowest occupied molecular orbital (LUMO) (d) of a C60 molecule.
Figure 2. Relaxed structure (a), total density of state (DOS) (b), highest occupied molecular orbital (HOMO) (c), and lowest occupied molecular orbital (LUMO) (d) of a C60 molecule.
Nanomaterials 09 01484 g002
Figure 3. DOS plots for the bulk structures of Cu (a), Ag (b) and Au (c). Dotted lines correspond to the Fermi energy.
Figure 3. DOS plots for the bulk structures of Cu (a), Ag (b) and Au (c). Dotted lines correspond to the Fermi energy.
Nanomaterials 09 01484 g003
Figure 4. (ac) Total DOSs for Cu, Ag and Au within a C60 molecule, respectively; (df) atomic doses for Cu, Ag and Au within a C60 molecule, respectively; and (gi) charge density distributions around the encapsulated metal atoms for Cu, Ag and Au within a C60 molecule, respectively.
Figure 4. (ac) Total DOSs for Cu, Ag and Au within a C60 molecule, respectively; (df) atomic doses for Cu, Ag and Au within a C60 molecule, respectively; and (gi) charge density distributions around the encapsulated metal atoms for Cu, Ag and Au within a C60 molecule, respectively.
Nanomaterials 09 01484 g004
Figure 5. Optimised structures of Cu (a), Ag (b) and Au (b) adsorbed on the surface of C60.
Figure 5. Optimised structures of Cu (a), Ag (b) and Au (b) adsorbed on the surface of C60.
Nanomaterials 09 01484 g005
Figure 6. (ac) Total DOSs for Cu, Ag and Au, respectively; (df) atomic doses for Cu, Ag and Au, respectively; and (gi) charge density distributions around the adsorbed metal atoms for Cu, Ag and Au, respectively.
Figure 6. (ac) Total DOSs for Cu, Ag and Au, respectively; (df) atomic doses for Cu, Ag and Au, respectively; and (gi) charge density distributions around the adsorbed metal atoms for Cu, Ag and Au, respectively.
Nanomaterials 09 01484 g006
Figure 7. (a) Relaxed structure of a defective C60 molecule (b) bond distances around the defect (c) DOS and (d) cross sectional charge density plot.
Figure 7. (a) Relaxed structure of a defective C60 molecule (b) bond distances around the defect (c) DOS and (d) cross sectional charge density plot.
Nanomaterials 09 01484 g007
Figure 8. Relaxed structures and charge density plots of Cu (a), Ag (b) and Au (c) encapsulated within a defective C60 molecule.
Figure 8. Relaxed structures and charge density plots of Cu (a), Ag (b) and Au (c) encapsulated within a defective C60 molecule.
Nanomaterials 09 01484 g008
Figure 9. Relaxed structures and charge density plots of Cu (a), Ag (b) and Au (c) adsorbed on the surface of defective C60 molecule.
Figure 9. Relaxed structures and charge density plots of Cu (a), Ag (b) and Au (c) adsorbed on the surface of defective C60 molecule.
Nanomaterials 09 01484 g009
Table 1. Calculated and experimental lattice constants and cohesive energies of bulk coinage metals.
Table 1. Calculated and experimental lattice constants and cohesive energies of bulk coinage metals.
Parameter
Lattice Constant (Å)Cohesive Energy (eV)
CuAgAuCuAgAu
Proposed method3.574.094.123.992.973.63
Experiment3.59 [47]4.06 [47]4.06 [47]3.48 [48]2.94 [48]3.81 [48]
Other theory3.501–3.686 [48]4.046–4.321 [48]4.084–4.112 [48]2.54–4.42 [48]1.87–3.60 [48]2.23–3.86 [48]
Table 2. Encapsulation energies of single coinage metal atoms incorporating the C60 molecule, Bader charges on metal atoms and magnetic moments of the resultant composites.
Table 2. Encapsulation energies of single coinage metal atoms incorporating the C60 molecule, Bader charges on metal atoms and magnetic moments of the resultant composites.
SystemAtomic Radius (Å)Encapsulation Energy (eV)Bader Charge |e|Magnetic Moment (µ)
Empirical [50]Calculated [51]DFTDFT + DDFTDFT + DDFTDFT + D
Cu@C601.351.450.13 −0.58+0.1675+0.16721.00001.0000
Ag@C601.601.650.32 −0.36+0.1704+0.17010.99830.9983
Au@C601.351.740.23 −0.56+0.0956+0.09540.99850.9985
Table 3. Initial and final configurations of coinage metal atoms adsorbed by the outer surface of the C60.
Table 3. Initial and final configurations of coinage metal atoms adsorbed by the outer surface of the C60.
Initial ConfigurationFinal Configuration and Relative Adsorption Energies (eV)
CuAgAu
HC (0.00)H (0.33)H (0.67)
PP (0.47)P (0.30)P (0.69)
6666 (0.12)C (0.00)C (0.00)
6565 (0.02)65 (0.07)C (0.00)
CC (0.00)C (0.00)C (0.00)
Table 4. Adsorption energies of atoms interacting with the surface of C60, Bader charge on the metal atoms and magnetic moments of the composites.
Table 4. Adsorption energies of atoms interacting with the surface of C60, Bader charge on the metal atoms and magnetic moments of the composites.
SystemAdsorption Energy (eV)Bader Charge |e|Magnetic Moment (µ)
Cu_C60−0.98 +0.300.9711
Ag_C60−0.50 +0.240.9329
Au_C60−0.89 +0.320.9778
Table 5. Encapsulation energies, Bader charge on the metal atoms and magnetic moments of the endohedral composites formed between the defective C60 molecule and metals.
Table 5. Encapsulation energies, Bader charge on the metal atoms and magnetic moments of the endohedral composites formed between the defective C60 molecule and metals.
SystemEncapsulation Energy (eV)Bader Charge |e|Magnetic Moment (µ)
Cu@C60_defe−0.56 +0.271.000
Ag@C60_defe−0.43 +0.491.000
Au@C60_defe−0.51 +0.311.000
Table 6. Encapsulation energies, Bader charge on the metal atoms and magnetic moments of the composites formed between the surface of defective C60 molecule and metals.
Table 6. Encapsulation energies, Bader charge on the metal atoms and magnetic moments of the composites formed between the surface of defective C60 molecule and metals.
SystemEncapsulation Energy (eV)Bader Charge |e|Magnetic Moment (µ)
Cu_C60_defe−0.61+0.700.8896
Ag_C60_defe−0.41+0.620.9012
Au_C60_defe−0.48+0.510.9414

Share and Cite

MDPI and ACS Style

Kuganathan, N.; Srikaran, R.; Chroneos, A. Stability of Coinage Metals Interacting with C60. Nanomaterials 2019, 9, 1484. https://doi.org/10.3390/nano9101484

AMA Style

Kuganathan N, Srikaran R, Chroneos A. Stability of Coinage Metals Interacting with C60. Nanomaterials. 2019; 9(10):1484. https://doi.org/10.3390/nano9101484

Chicago/Turabian Style

Kuganathan, Navaratnarajah, Ratnasothy Srikaran, and Alexander Chroneos. 2019. "Stability of Coinage Metals Interacting with C60" Nanomaterials 9, no. 10: 1484. https://doi.org/10.3390/nano9101484

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