Quantum Chemical Studies of Endofullerenes (M@C60) Where M = H2O, Li+, Na+, K+, Be2+, Mg2+, and Ca2+

Abstract

Quantum chemical calculations were performed to investigate the structural and electronic properties of seven endofullerenes. The interaction energies indicated that all of the chemical species are stable inside the fullerene for each complex. The ionization potential and electron affinity values suggest that the endofullerenes consisting of alkaline earth ions are the most reactive and that the dipole moment decreased according to the following order: Be2+@C60 (4.75) > Mg2+@C60 (3.14) > Ca2+@C60 (2.24) > Li+@C60 (1.26) > Na+@C60 (0.76) > H2O@C60 (0.24) > K+@C60 (0.00 Debye). These results imply that the solubility of endofullerenes in a polar solvent may increase from H2O@C60 to Be2+@C60. The energetic gaps indicate that Be2+@C60 and Mg2+@C60 possess a higher electrical conductivity, and the UV spectra show a shift in the bands to the visible light region. The results of this work may be useful for the development of new endofullerenes.

Share and Cite:

Oliveira, O. and Gonçalves, A. (2014) Quantum Chemical Studies of Endofullerenes (M@C60) Where M = H2O, Li+, Na+, K+, Be2+, Mg2+, and Ca2+. Computational Chemistry, 2, 51-58. doi: 10.4236/cc.2014.24007.

1. Introduction

Fullerene (C60), which was discovered in 1985 by Kroto et al., is the third allotropic form of carbon [1] . The most well-known structure of C60 is the Buckminsterfullerene, which consists of 20 hexagonal rings and 12 pentagonal rings that form a geometric structure with 32 faces, which form a closed cage. Each carbon atom is bonded to three others with sp2 hybridization, and C60 tends to react readily with electron rich species due its poor electron delocalization. Since the discovery of C60 [1] , the literature has reported the development of several fullerene derivatives based on the functionalization of the C60 cage aiming to produce new properties and/or improve their existing properties. Other important studies involve the use of C60 derivatives in biological systems [2] -[4] , solar cells [5] -[8] and superconductors [9] -[11] . Another class of C60 derivatives has been achieved by the encapsulation of small molecules, metals and ions inside the C60 cage. This type of system is known as an endofullerene (nomenclature proposed by Chai et al.) [12] . This representation is referred to as A@Cn, where A and n are the chemical species and the number of carbon atoms, respectively. Among the endofullerenes, those based on C60 have attracted much attention due their potential applications in superconductivity and materials science [13] -[15] . The first evidence of metal trapped inside C60 was obtained by Heath and collaborators [16] , who reported the formation of the La@C60 complex. In the same laboratory, an air-stable film of La@C60 was obtained [12] . Li+@C60 was the first endohedral to be isolated, characterized and its crystal structure elucidated [17] [18] . Other non-metal C60 endofullerenes have been experimentally determined including H2@C60 [19] [20] , N2@C60 [21] , noble gas [22] [23] , and, more recently, H2O@C60 [24] -[26] .

C60 endofullerenes can be obtained using three different experimental methods [27] including 1) the vaporization of graphite, 2) the insertion of the atoms through the fullerene cages, and 3) chemical routes that involve opening the fullerene cage. However, this method involves the use of extreme conditions, such as high temperature and treatment with high pressure. The increased energy expenditure results in endofullerene synthesis having high costs. Theoretical methods based on quantum chemistry are useful for predicting important properties that, in general, confirm the experimental data. Several theoretical studies of C60 endofullerenes have been reported. Among these studies, quantum chemistry has been employed to study 1) the structural and optical properties of endofullerenes [28] , 2) the mechanism of alkali-metal insertion into C60 [29] , 3) electronic structures [30] [31] , 4) the charge transfer from the metal to C60 [32] [33] , 5) the insertion of Be into C60 [34] , and 6) the photoionization cross section in Mg2+@C60 [35] . Although some alkali and alkaline earth ions have been studied using quantum chemical methods [28] -[35] , the reactivity parameters of the current endofullerenes have not been predicted. However, some properties are very important in the chemical synthesis. Therefore, the goal of the current work was to investigate the electronic and structural properties of M@C60 (where M = H2O, Li+, Na+, K+, Be2+, Mg2+, and Ca2+) using quantum chemical methods. We have considered H2O@C60 and pure C60 to be reference systems when comparing the properties of a non-metallic endofullerene with the metallic ions complexes. Here, we are interested in predicting the reactivity parameters, electrical conductivity, absorption spectra, polarization and charge transfer process.

2. Computational Details

The C60 structure was retrieved from the Library of 3-D Molecular Structures [http://www.nyu.edu/pages/mathmol/library/]. The endofullerenes were initially constructed by inserting the chemical species into the center of the fullerene using PyMOL [http://www.pymol.org/]. In sequential calculations, all of the structures were initially optimized using semi-empirical PM7 Hamiltonian [36] , as implemented in the MOPAC package [37] . The final structures from these calculations were re-optimized using the GAMESS-US software [38] using DFT/B3LYP hybrid functional [39] [40] with the 6-31G(2d,2p) basis set. All of the stationary points were characterized as a point of minimum energy using the harmonic vibrational and ZPE correction calculations with the 6-31G(2d,2p) basis set. The absorption spectra were calculated using the TDDFT method with the 6-31G basis set, as implemented in the GAMESS-US software [38] . This basis set was used to reduce the high computational cost of the TDDFT calculations, and the polarizable continuum model (PCM) [41] of solvation was employed to account for the solvent (toluene) effects.

3. Results and Discussion

At the end of the optimizations, which were conducted using different methodologies, the spatial structures were aligned, and the RMSD (root mean square deviation) was calculated based on the DFT optimized structure, which was used as a reference. Surprisingly, the RMSD values were lower than 0.001 Å for all of the endofullerenes except Be2+@C60 and Mg2+@C60, for which the RMSD was 0.01 Å. These small values indicate that the endofullerene structures obtained using the semi-empirical PM7 and DFT methods are very similar. Therefore, these methods possessed similar accuracies for the prediction of the current endofullerene structures. However, all of the reported results have been extracted from the DFT optimized structures (Figure 1).

The results in Figure 1 indicate that in all of the endofullerenes except H2O@C60 and K+@C60, the ions are

Figure 1. DFT/B3LYP/6-31G** optimized structures. The ion size is proportional to the van der Waals radius. The carbon atoms are shown in green.

located in an off-center position in the C60. In addition, these ions are near the six-membered ring, with a distance of 2.3, 2.8, 1.8 and 2.6 Å for Li+, Na+, Be2+ and Ca2+ ions, respectively. The Li+...C distance in the crystallographic structure (2.3 Å) [17] and our optimized molecule are the same, which demonstrates the accuracy of the current theoretical protocols. In K+@C60, the K+ ion is located in the center of the C60 cage, which is in agreement with the results from previous calculations [42] . This behavior can be attributed to the high ionic size of K+ compared with the other studied ions. For the optimized H2O@C60 complex, the shortest H...C and O...C distances are 2.70 and 3.45 Å, respectively, which is consistent with the crystallographic structure [26] , where the H...C and O...C interatomic distances are 2.60 and 3.55 Å, respectively. In addition, we performed an optimization with a single water molecule using the same theoretical level to compare the various properties obtained with the water molecule encapsulated in C60. In this calculation, the H-O bonds had the same distance (0.96 Å). However, the H-O-H angle was different (i.e., H2O@C60 = 103.72˚ and pure water = 104.13˚). A decrease of 0.4% in this angle was most likely due to electronic repulsion between the hydrogen and carbon atoms. The endofullerene stability was based on the total energy (in hartree) obtained by DFT optimizations. The stability of the endofullerenes increased in the following order: C60 (−2284.8) > Li+@C60 (−2292.2) > Be2+@C60 (−2298.9) > H2O@C60 (−2361.2) > Na+@C60 (−2446.9) > Mg2+@C60 (−2484.2) > K+@C60 (−2884.5) > Ca2+@C60 (−2961.8).

In all of the optimized structures, the encapsulation of the chemical species is favorable (negative values of ∆E). Table 1 summarizes some of the electronic properties obtained from the calculations.

According to Koopmans’ theorem, the ionization potential (IP) of a molecule is the highest orbital molecular orbital (HOMO) energy (εHOMO) multiplied by −1, and the lowest unoccupied molecular orbital (LUMO) energy (εLUMO) multiplied by −1 represents the electron affinity (EA) or the capacity of a molecule to accept electrons. In a simplistic reaction model, the LUMO would be expected to accept electrons from the HOMO of another molecule. The results in Table 1 indicate that the endofullerenes containing ions in the same family have a similar IP, with values of ~9.20 and ~12.36 eV for alkali and alkaline earth ions, respectively. Therefore, the size of ionic radius and the IP of the ions in the same family do not effectively contribute to the IP of C60 endofullerenes. However, the IP of endofullerenes increases by ~31% when alkali ions are replaced by alkaline earth ions. This increase is due to the high IP values of the alkaline earth ions compared with the alkali ions. Interestingly, the presence of the water molecule in the C60 decreased its IP, which indicated that the C60 endofullerene is more electron donor than the pure C60. In contrast, an increase in the IP was observed for the metal endofullerenes because they are more electron acceptor than C60. Therefore, in both cases, the ions transfer their ionic character to C60 in the encapsulation process. The same behavior was observed for the electron affinity (LUMO orbital energy), as shown in Table 1. The EA is more negative for all of the C60 endofullerenes studied than it is for pure C60, which indicated that the endofullerenes are more electron accepting compared with pure fullerene. Other chemical properties that can be estimated from the εLUMO and εHOMO include the chemical potential, µ ((εLUMO + εHOMO)/2), and the chemical hardness, η((εLUMO − εHOMO)/2). µ and η measure the tendency of an electron to escape from a particular field [43] and the resistance to modification of the electronic density [44] , respectively. Therefore, the low values of µ and η for the alkaline earth ion C60 endofullerenes indicated that these

Table 1. Properties obtained from quantum chemical calculations at the DFT/B3LYP/6-31G** level.

a in kcal/mol; bChelpG charge of encapsulated chemical species.

complexes are more resistant to donating an electron and that their electronic density is more easily modified compared with the other endofullerenes. In other words, the endofullerenes based on alkaline earth ions are more reactive than are the other endofullerenes studied. Among the M@C60 complexes, Mg2+@C60 exhibited the lowest values of µ (−11.66 eV) and η (0.70 eV), which is expected to result in high reactivity. The degree of fullerene polarization was characterized by the dipole moment. Except for K+@C60, all of the studied chemical species increased the C60 polarizability, and the dipole moment values increased with the decreasing binding energies (∆E) (Table 1). The K+ ion did not increase the dipole of C60 due to the large distance between the carbon atoms and this ion. The dipole moment for H2O@C60 was 0.34 D. Therefore, the water molecule transferred its partial dipole to C60. The calculated dipole moment for pure water obtained using the same theoretical level as that employed for all of the endofullerenes was 2.04 D. By comparing the dipole moment of pure H2O and H2O@C60, a decrease of 83% was observed, which is in agreement with the ab initio results reported by Shameema [45] . However, Kurotobi and Murata [24] predicted an increase in the dipole moment of 0.05% in H2O@C60 relative to pure water using the MX06-2X functional. This discrepancy may be due to the different functionals used. The dipole moment of the endofullerenes decreased in the following order: Be2+@C60 (4.75 D) > Mg2+@C60 (3.14 D) > Ca2+@C60 (2.24 D) > Li+@C60 (1.26 D) > Na+@C60 (0.76 D) > H2O@C60 (0.24 D) > K+@C60 (0.00 D). This sequence indicates that the solubility of endofullerene in water or other polar solvents may increase from H2O@C60 to Be2+@C60. Therefore, the presence of these chemical species, except for K+, in the C60 cage can improve its solubility. The dipole moment observed for these fullerenes was due to the charge transfer from the chemical species to the fullerene (Table 1). Except for water, all of the chemical species positively increased the fullerene charge. For single water, the ChelpG charge of the O and H atoms were −0.72 and 0.36, respectively. In H2O@C60, the charge of the oxygen increased to −0.46, and the charge of the hydrogen atoms decreased to 0.33 and 0.32. Therefore, a portion of the water charge was transferable to C60, which resulted in a negative charge of −0.19 (Table 1). For all of the systems, the carbon atoms nearest the ions acquired a negative charge, while the positive charge was distributed over the C60 surface that was far from the metal ions. In general, the charge transfer from the cation to C60 increased as the binding energy decreased due to cation-pi electronic interactions. Therefore, Be2+ transferred 82% of its charge to the fullerene, and this endofullerene possessed the lowest binding energy (−11.96 kcal/mol).

Electronic conductivity is another interesting and important property of the new materials. This property can be estimated from the difference between the calculated HOMO and LUMO energies using Equation (1).

(1)

where σ is the electric conductivity, Eg is the band gap energy, kb is Boltzmann’s constant and T is the temperature. This equation indicates that a higher electric conductivity is due to a decrease in Eg. Figure 2 shows the HOMO and LUMO energies and their surface representations obtained from DFT calculations.

C60 acts as a semiconductor with an experimental band gap of 2.3 - 2.7 eV [46] -[49] . Therefore, the value (2.90 eV) that we obtained using the approximation |HOMO − LUMO| (Figure 2) is close to the experimental data. It is interesting to note that for all of the endofullerenes, the calculated energy gap is less than that of pure C60. Based on Equation (1) and the low band gap energy of Be2+@C60 (1.98 eV) and Mg2+@C60 (1.38 eV), we suggest that these endofullerenes are more conductive than C60 and the other studied endofullerenes.

Another important property in the development of new materials is the capacity of a material to absorb in the visible region. Figure 3 shows the UV spectra of all of the studied compounds.

Figure 2. HOMO and LUMO orbital energies obtained at the B3LYP/6-31G(2d,2p) level.

Figure 3. UV spectra obtained at the B3LYP/6-31G level. Inset shows the zoom of the region (250 - 500 nm, 0 - 0.38).

According to Prylutskyy [50] , fullerene in toluene solvent possesses two intense broad absorption bands (i.e., 269 and 335 nm) and allowed transitions. In the current work, the band of major interest is near the UV-visible region (335 nm) because we believe that the presence of ions in C60 can shift this band to the visible region. In our calculations, we observed a band centered at 310 nm for a single C60. Because our calculations were performed using PCM [41] with average basis sets, the calculated UV spectra are in reasonable agreement with the experimental data. However, the band at 310 nm for pure C60 is in agreement with previous theoretical calculations [51] [52] . In Figure 3, the absorption and band intensity of H2O@C60 and K+@C60 are similar to those of pure C60, which is due to the low interaction energy between these species and C60 (Table 1). For Li+@C60 and Na+@C60, the same wavelength (308 nm) with a relative intensity of 0.34 and a small shoulder at 343 nm was also observed. Interestingly, two bands at 343 (small shoulder) and 350 nm (large shoulder) were observed for Ca2+@C60 and Be2+@C60, respectively. However, the largest shoulder with a band centered at 356 nm with the lowest relative intensity (0.09) was observed for Mg2+@C60. This result reinforces the hypothesis that the presence of Mg2+ in C60 can shift the absorption spectra to the visible region (Figure 3).

4. Conclusion

Chemical species trapped in fullerenes comprise a new class of molecules with interesting properties. Therefore, in the current work, we employed quantum chemical methods to study seven endofullerenes (i.e., H2O@C60, Li+@C60, Na+@C60, K+@C60, Be2+@C60, Mg2+@C60 and Ca2+@C60). Our results indicate that all of the endoful- lerenes are more stable than pure C60. The total energy (in hartree) calculated using DFT optimizations is as follows: C60 (−2284.8) > Li+@C60 (−2292.2) > Be2+@C60 (−2298.9) > H2O@C60 (−2361.2) > Na+@C60 (−2446.9) > Mg2+@C60 (−2484.2) > K+@C60 (−2884.5) > Ca2+@C60 (−2961.8). The reactivity parameters (i.e., ionization potential and electron affinity) indicate that the alkali earth ions trapped in C60 are the most reactive. In our calculations, charge transfer was observed from the ions to the fullerene. Therefore, except for K+@C60, the dipole moment increased in the following order: Be2+@C60 (4.75 Debye) > Mg2+@C60 (3.14 D) > Ca2+@C60 (2.24 D) > Li+@C60 (1.26 D) > Na+@C60 (0.76 D) > H2O@C60 (0.24 D) > K+@C60 (0.00 D). Therefore, the solubility of the endofullerenes in polar solvents is expected to follow the same trend. Based on the HOMO and LUMO energy difference, Be2+@C60 and Mg2+@C60 have the highest electrical conductivity with band gap energies of 1.98 and 1.38 eV, respectively. The UV spectra obtained is in reasonable agreement with the experimental data. The C60 band centered at 310 nm shifted to 343, 350 and 356 nm for Ca2+@C60, Be2+@C60 and Mg2+@C60, respectively, which confirms the previous theoretical results.

Acknowledgements

We wish to thank our institute for financial support for this study.

NOTES

*Corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

[1] Kroto, H.W., Heath, J.R., O’Brien, S.C., Curl, R.F. and Smalley, R.E. (2003) C60: Buckminsterfullerene. Nature, 318, 162-163. http://dx.doi.org/10.1038/318162a0
[2] Bosi, S., Ros, T.D., Spalluto, G. and Prato, M. (2003) Fullerene Derivatives: An Attractive Tool for Biological Applications. European Journal of Medicinal Chemistry, 38, 913-923.
http://dx.doi.org/10.1016/j.ejmech.2003.09.005
[3] Bakry, R., Vallant, R.M., Najam-ul-Haq, M., Rainer, M., Szabo, Z., Huck, C.W. and Bonn, G.K. (2007) Medicinal Applications of Fullerenes. International Journal of Nanomedicine, 2, 639-649.
[4] Anilkumar, P., Lu, F., Cao, L., Luo, P.G., Liu, J.H., Sahu, S., Tackett, K.N., Wang, Y. and Sun, Y.P. (2011) Fullerenes for Applications in Biology and Medicine. Current Medicinal Chemistry, 18, 2045-2059. http://dx.doi.org/10.2174/092986711795656225
[5] Blom, P.W.M., Mihailetchi, V.D., Koster, L.J.A. and Markov, D.E. (2007) Device Physics of Polymer: Fullerene Bulk Heterojunction Solar Cells. Advanced Materials, 19, 1551-1566.
http://dx.doi.org/10.1002/adma.200601093
[6] Deibel, C. and Dyakonov, V. (2010) Polymer-Fullerene Bulk Heterojunction Solar Cells. Reports on Progress in Physics, 73, Article ID: 096401. http://dx.doi.org/10.1088/0034-4885/73/9/096401
[7] Varotto, A., Treat, N.D., Jo, J., Shuttle, C.G., Batara, N.A., Brunetti, F.G., Seo, J.H., Chabinyc, M.L., Hawker, C.J., Heeger, A.J. and Wudl, F. (2011)1,4-Fullerene Derivatives: Tuning the Properties of the Electron Transporting Layer in Bulk-Heterojunction Solar Cells. Angewandte Chemie, 50, 5166-5169.
[8] Liu, B., Png, R.Q., Zhao, L.H., Chua, L.L., Friend, R.H. and Ho, P.K. (2012) High Internal Quantum Efficiency in Fullerene Solar Cells Based on Crosslinked Polymer Donor Networks. Nature Communications, 3, 1321. http://dx.doi.org/10.1038/ncomms2211
[9] Tanigaki, K., Ebbesen, T.W., Saito, S., Mizuki, J., Tsai, J.S., Kubo, Y. and Kuroshima, S. (1991) Superconductivity at 33 K in CsxRbyC60. Nature, 352, 222. http://dx.doi.org/10.1038/352222a0
[10] Buntar, V. and Weber, H.W. (1996) Magnetic Properties of Fullerene Superconductors. Superconductor Science and Technology, 9, 599-615. http://dx.doi.org/10.1088/0953-2048/9/8/001
[11] Hott, R., Kleiner, R., Wolf, T. and Zwicknagl, G. (2013) Review on Superconducting Materials. Wiley-VCH, Hoboken.
[12] Chai, Y., Guo, T., Jin, C.M., Haufler, R.E., Chibante, L.P.F., Fure, J., Wang, L.H., Alford, J.M. and Smalley, R.E. (1991) Fullerenes with Metals Inside. Journal of Physical Chemistry, 95, 7564-7568.
http://dx.doi.org/10.1021/j100173a002
[13] Bethune, D.S., Johnson, R.D., Salem, J.R., Vries, M.S. and Yannoni, C.S. (1993) Atoms in Carbon Cages: The Structure and Properties of Endohedral Fullerenes. Nature, 366, 123-128.
http://dx.doi.org/10.1038/366123a0
[14] Hirata, T., Hatakeyama, R., Mieno, T. and Sato, N. (1996) Production and Control of K-C60 Plasma for Material Processing. Journal of Vacuum Science & Technology A, 14, 615-619.
http://dx.doi.org/10.1116/1.580154
[15] Chaur, M.N., Melin, F., Ortiz, A.L. and Echegoyen, L. (2009) Chemical, Electrochemical, and Structural Properties of Endohedral Metallofullerenes. Angewandte Chemie International Edition, 48, 7514-7538. http://dx.doi.org/10.1002/anie.200901746
[16] Heath, J.R., O’Brien, S.C., Zhang, Q., Liu, Y., Curl, R.F., Tittel, F.K. and Smalley, R.E. (1985) Lanthanum Complexes of Spheroidal Carbon Shells. Journal of the American Chemical Society, 107, 7779-7780. http://dx.doi.org/10.1021/ja00311a102
[17] Aoyagi, S., Nishibori, E., Sawa, H., Sugimoto, K., Takata, M., Miyata, Y., Kitaura, R., Shinohara, H., Okada, H., Sakai, T., et al. (2010) A Layered Ionic Crystals of Polar Li@C60 Superatoms. Nature Chemistry, 2, 678-683. http://dx.doi.org/10.1038/nchem.698
[18] Okada, H., Komuro, T., Sakai, T., Matsuo, Y., Ono, Y., Omote, K., Yokoo, K., Kawachi, K., Kasama, Y., Ono, S., et al. (2012) Preparation of Endohedral Fullerene Containing Lithium (Li@C60) and Isolation as Pure Hexafluorophosphate Salt ([Li+@C60][PF6-]). RSC Advances, 2, 10624-10631.
http://dx.doi.org/10.1039/c2ra21244g
[19] Murata, Y., Murata, M. and Komatsu, K. (2003) 100% Encapsulation of a Hydrogen Molecule into an Open-Cage Fullerene Derivative and Gas-Phase Generation of H2@C60. Journal of the American Chemical Society, 125, 7152-7153. http://dx.doi.org/10.1021/ja0354162
[20] Komatsu, K., Murata, M. and Murata, Y. (2005) Encapsulation of Molecular Hydrogen in Fullerene C60 by Organic Synthesis. Science, 307, 238-240. http://dx.doi.org/10.1126/science.1106185
[21] Ito, S., Shimotani, H., Takagi, H. and Dragoe, N. (2008) On the Synthesis Conditions of N and N2 Endohedral Fullerenes. Fullerenes, Nanotubes and Carbon Nanostructures, 16, 206-213.
[22] Sauders, M., Cross, R.J., Jimenez-Vazquez, H.A., Shimshi, R. and Khong, A. (1996) Noble Gas Atoms inside Fullerences. Science, 271, 1693-1697. http://dx.doi.org/10.1126/science.271.5256.1693
[23] Peng, R.F., Chu, S.J., Hyang, Y.M., Yu, H.J., Wang, T.S., Jin, B., Fu, Y.B. and Wang, C.R. (2009) Preparation of He@C60 and He2@C60 by an Explosive Method. Journal of Materials Chemistry, 19, 3602-3605. http://dx.doi.org/10.1039/b904234b
[24] Kurotobi, K. and Murata, Y. (2011) A Single Molecule of Water Encapsulated in Fullerene C60. Science, 333, 613-616.
[25] Beduz, C., Carravetta, M., Chen, J.C., Consistrè, M., Denning, M., Frunzi, M., et al. (2012) Quantum Rotation of Ortho and Para-Water Encapsulated in a Fullerene Cage. Proceedings of the National Academy of Sciences of the United States of America, 109, 12894-12898.
http://dx.doi.org/10.1073/pnas.1210790109
[26] Aoyagi, S., Hoshino, N., Akutagawa, T., Sado, Y., Kitaura, R., Shinohrara, H., Sugimoto, K., Zhang, R. and Murata, Y. (2014) A Cubic Dipole Lattice of Water Molecules Trapped inside Carbon Cages. Chemical Communications, 50, 524-526. http://dx.doi.org/10.1039/c3cc46683c
[27] Popov, A.A., Yang, S. and Dunsch, L. (2013) Endohedral Fullerenes. Chemical Reviews, 113, 5989-6113. http://dx.doi.org/10.1021/cr300297r
[28] Noguchi, Y., Sugino, O., Okada, H. and Matsuo, Y. (2013) First-Principles Investigation on Structural and Optical Properties of M+@C60 (Where M = H, Li, Na, and K). Journal of Physical Chemistry C, 117, 15362-15368. http://dx.doi.org/10.1021/jp4041259
[29] Malani, H. and Zhang, D. (2013) Theoretical Insight for the Metal Insertion Pathway of Endohedral Alkali Metal Fullerenes. Journal of Physical Chemistry A, 117, 3521-3528.
http://dx.doi.org/10.1021/jp4007697
[30] Cioslowski, J. (1991) Endohedral Chemistry: Electronic Structures of Molecules Trapped inside the C60 Cage. Journal of the American Chemical Society, 113, 4139-4141.
http://dx.doi.org/10.1021/ja00011a013
[31] Cioslowski, J. and Fleischmann, E.D. (1991) Endohedral Complexes: Atoms and Ions inside the C60 Cage. Journal of Chemical Physics, 94, 3730-3734. http://dx.doi.org/10.1063/1.459744
[32] Hira, A.S. and Ray, A.K. (1995) Interaction Sites of a Na+ Ion and a Na Atom with a C60 Molecule. Physical Review A, 52, 141-148. http://dx.doi.org/10.1103/PhysRevA.52.141
[33] Santos, J.D., Longo, E., Banja, M.E., Espinoza, V.A.A., Flores, J.V. and Taft, C.A. (2005) Semi-Empirical Studies of Alkaline Metals-Fullerene MxC60, M@C60 Interactions. Journal of Molecular Structure: THEOCHEM, 713, 161-169. http://dx.doi.org/10.1016/j.theochem.2004.08.055
[34] Ohtsuki, T., Masumoto, K., Ohno, K., Maruyma, Y., Kawazoe, Y., Sueki, K. and Kikuchi, K. (1996) Insertion of Be Atoms in C60 Fullerene Cages: Be@C60. Physical Review Letters, 77, 3522-3524.
http://dx.doi.org/10.1103/PhysRevLett.77.3522
[35] Lyras, A. and Bachau, H. (2005) Electronic Correlation Effects in a Model of Endohedral Mg (Mg@C60). Journal of Physics B: Atomic, Molecular and Optical Physics, 38, 1119-1131.
http://dx.doi.org/10.1088/0953-4075/38/8/004
[36] Stewart, J.J.P. (2013) Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters. Journal of Molecular Modeling, 19, 1-32. http://dx.doi.org/10.1007/s00894-012-1667-x
[37] Stewart, J.J.P. (2012) MOPAC2012, Stewart Computational Chemistry. Colorado Springs, CO.
HTTP://OpenMOPAC.net
[38] Schmidt, M.W., Baldridge, K.K., Boatz, J.A., Elbert, S.T., Gordon, M.S., Jensen, J.H., Koseki, S., Matsunaga, N., Nguyen, K.A., Su, S., Windus, T.L., Dupuis, M. and Montgomery, J.A. (1993) General Atomic and Molecular Electronic Structure System. Journal of Computational Chemistry, 14, 1347-1363. http://dx.doi.org/10.1002/jcc.540141112
[39] Lee, C., Yang, W. and Parr, R.G. (1998) Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Physical Review B, 37, 785-789.
http://dx.doi.org/10.1103/PhysRevB.37.785
[40] Becke, A.D. (1993) Density Functional Thermochemistry. III. The Role of Exact Exchange. Journal of Chemical Physics, 98, 5648-5652. http://dx.doi.org/10.1063/1.464913
[41] Tomasi, J., Mennucci, B. and Cammi, R. (2005) Quantum Mechanical Continuum Solvation Models. Chemical Reviews, 105, 2999-3094. http://dx.doi.org/10.1021/cr9904009
[42] Dunlap, B.I., Ballester, J.L. and Schmidt, P.P. (1992) Interactions between Fullerene C60 and Endohedral Alkali Atoms. Journal of Physical Chemistry, 96, 9781-9787.
http://dx.doi.org/10.1021/j100203a038
[43] Parr, R.G., Donnelly, R.A., Levy, M. and Palke, W.E. (1978) Electronegativity: The Density Functional Viewpoint. Journal of Chemical Physics, 68, 3801-3807. http://dx.doi.org/10.1063/1.436185
[44] Parr, R.G. and Pearson, R.G. (1983) Absolute Hardness: Companion Parameter to Absolute Electronegativity. Journal of the American Chemical Society, 105, 7512-7516.
http://dx.doi.org/10.1021/ja00364a005
[45] Shameema, O., Ramachandran, C.N. and Sathyamurthy, N. (2006) Blue Shift in X-H Stretching Frequency of Molecules Due to Confinement. Journal of Physical Chemistry A, 110, 2-4.
http://dx.doi.org/10.1021/jp056027s
[46] Benning, P.J., Martins, J.L., Weaver, J.H., Chibante, L.P.F. and Smalley, R.E. (1991) Electronic Structure of KxC60: Insulating, Metallic, and Superconducting Character. Science, 252, 1417-1419.
http://dx.doi.org/10.1126/science.252.5011.1417
[47] Takahashi, T., Susuzi, S., Morikawa, T., Katayama-Yoshida, H., Hasegawa, S., Inokuchi, H., Seki, K., Kikuchi, K., Suzuki, S., Ikemoto, K. and Ashiba, Y. (1992) Pseudo-Gap at the Fermi Level in K3C60 Observed by Photoemission and Inverse Photoemission. Physical Review Letters, 68, 1232.
http://dx.doi.org/10.1103/PhysRevLett.68.1232
[48] Lof, R.W., van Veendaal, M.A., Koopmans, B., Jonkman, H.T. and Sawatzky, G.A. (1992) Band Gap, Excitons, and Coulomb Interaction in Solid C60. Physical Review Letters, 68, 3924.
http://dx.doi.org/10.1103/PhysRevLett.68.3924
[49] Weaver, J.H. (1992) Electronic Structures of C60, C70 and the Fullerides: Photoemission and Inverse Photoemission Studies. Journal of Physics and Chemistry of Solids, 53, 1433-1447.
http://dx.doi.org/10.1016/0022-3697(92)90237-8
[50] Prylutskyy, Y.I., Durov, S.S., Bulavin, L.A., Adamenko, I., Moroz, K.O., Graja, A., Bogucki, A. and Scharff, P. (2001) C 1s Ionisation Potential and Energy Referencing for Solid C60 Films on Metal Surfaces. Fullerene Science and Technology, 9, 167-174. http://dx.doi.org/10.1081/FST-100102964
[51] Braga, M., Larsson, S., Rosen, A. and Volosov, A. (1991) Electronic Transition in C60. On the Origin of the Strong Interstellar Absorption at 217nm. Astronomy and Astrophysics, 245, 232-238.
[52] Orlandi, G. and Negri, F. (2002) Electronic States and Transitions in C60 and C70 Fullerenes. Photochemical & Photobiological Sciences, 1, 289-308. http://dx.doi.org/10.1039/b200178k

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.