Next Article in Journal
Citrus-Derived Carbon Quantum Dots: Synthesis, Characterization, and Safety Evaluation in Zebrafish (Danio rerio) for Potential Biomedical and Nutritional Applications
Previous Article in Journal
Early Insights into AI and Machine Learning Applications in Hydrogel Microneedles: A Short Review
Previous Article in Special Issue
Evaluation of the Influence of Lorentz Forces on the Natural Frequencies of a Dual-Microcantilever Sensor for Ultralow Mass Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Brief Report

A Note on Computational Characterization of Dy@C82: Dopant for Solar Cells

1
Department of Chemistry and Biochemistry, University of Arizona, Tucson, AZ 85721-0041, USA
2
State Key Laboratory of Materials Processing and Die & Mould Technology, School of Material Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
3
Department of Physical and Macromolecular Chemistry, Faculty of Science, Charles University in Prague, Albertov 6, 128 43 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Micro 2025, 5(4), 49; https://doi.org/10.3390/micro5040049
Submission received: 8 August 2025 / Revised: 28 October 2025 / Accepted: 30 October 2025 / Published: 31 October 2025

Abstract

Dy@C82 is one of the metallofullerenes studied as dopants for improvements of stability and performance of solar cells. Calculations should help in formulating rules for selections of fullerene endohedrals for such new applications in photovoltaics. Structure, energetics, and relative equilibrium populations of two potential-energy-lowest IPR (isolated pentagon rule) isomers of Dy@C82 under high synthetic temperatures are calculated using the Gibbs energy based on molecular characteristics at the B3LYP/6-31G*∼SDD level. Dy@ C 2 v ( 9 ) -C82 and Dy@ C s ( 6 ) -C82 are calculated as 58 and 42%, respectively, of their equilibrium mixture at a synthetic temperature of 1000 K, in agreement with observations. The Dy@ C 2 v ( 9 ) -C82 species is found as lower in the potential energy by 1.77 kcal/mol compared to the Dy@ C s ( 6 ) -C82 isomer.

1. Introduction

Fullerenes and metallofullerenes have recently been considered as dopants for improvements of stability and performance of solar cells [1,2,3,4,5,6,7], for example, Er@C82 (the improvements could be related [5] to its hydrophobicity). Applications of various fullerene derivatives and tuning their optical and electronic characteristics in order to increase the light conversion efficiency in organic solar cells are surveyed in a recent review by Mumyatov and Troshin [7].
Dy@C82 represents another metallofullerene species studied for augmentation of photovoltaic systems. Yang et al. [1] treated the Langmuir–Blodgett films with Dy@C82 and reported enhanced quantum yields, up to nine-fold higher [2,3]. Actually, just two isomers of Dy@C82 are characterized by observations, namely [8,9,10,11,12,13,14,15,16] the species with the cage symmetries C 2 v ( 9 ) (major isomer) and C s ( 6 ) (minor isomer). Iida et al. concluded [8] the C 2 v cage symmetry for the major species while Takabayashi et al. arrived [9] at the C s cage symmetry with the minor isomer (more specifically, the cage is coded as C s ( c ) or C s ( 6 ) ). It should be mentioned that there are two cage-labeling conventions [17] applied in the literature for the altogether nine C82 IPR (isolated pentagon rule) fullerene cages. The older system refers to the symmetries of the carbon cages (actually to their highest topological symmetries, i.e., without possible symmetry reductions caused, for example, by the Jahn–Teller distortions): C 3 v ( a ) , C 3 v ( b ) , C 2 v , C 2 ( a ) , C 2 ( b ) , C 2 ( c ) , C s ( a ) , C s ( b ) , and C s ( c ) . The newer labeling system employs serial isomer-enumeration numbers [17] produced by the so-called spiral algorithm: C 3 v ( 7 ) , C 3 v ( 8 ) , C 2 v ( 9 ) , C 2 ( 3 ) , C 2 ( 1 ) , C 2 ( 5 ) , C s ( 2 ) , C s ( 4 ) , and C s ( 6 ) . Metal encapsulation into the C82 cages frequently leads to at least two isomeric endohedral species as known from both observations and calculations [17,18,19,20,21,22,23]. The present work is focused on the two Dy@C82 IPR isomers characterized in observations (the two isomers are the lowest structures in the potential energy). In particular, the temperature-dependent equilibrium relative isomeric populations are calculated for the two Dy@C82 isomers: Dy@ C 2 v ( 9 ) -C82 and Dy@ C s ( 6 ) -C82. Such calculated relative equilibrium populations are useful in interpretations of experimental data obtained for conditions close to the inter-isomeric thermodynamic equilibrium. In particular, the calculated characteristics should be helpful in formulations of rules for future rational screening of fullerene endohedrals for their applications in photovoltaics. The first step towards such selection rules is accumulation of calculated molecular characteristics of metallofullerenes already known [1,2,3,4,5,6,7] as applicable dopants for improvements of stability and performance of solar cells.
Let us mention that the stabilities of nanocarbons, and in particular metallofullerenes, are usually treated using just potential-energy terms (i.e., without evaluation of temperature developments of their populations). Nevertheless, there are several calculations [24,25,26,27,28,29,30,31,32,33] showing that the Gibbs energy should be considered instead, as the entropic part becomes increasingly important when temperature raises. In particular, a structure that is not the lowest in potential energy can still become the most populated species at high synthetic temperatures (used in nanocarbon preparations, for example, with production of metallofullerenes). Clearly enough, it is not possible to obtain such more complex relative-stability interplay only from the mere potential energies. Thus, calculations are performed in this study on the equilibrium relative populations of the two observed (and the potential-energy-lowest) Dy@C82 IPR isomers at higher temperatures, considering both the enthalpy and entropy parts of the Gibbs energy in order to predict their reliable equilibrium isomeric populations at high-temperature synthetic conditions.

2. Calculations

The molecular structures of the two Dy@C82 isomers were optimized with the density functional theory (DFT) approach, first using a combined atomic basis set—the standard 3-21G basis [34] for C atoms and SDD basis [35] with the SDD effective core potential on Dy atom (denoted here as 3-21G∼SDD). The applied DFT treatment combines Becke’s three-parameter functional [36] with the non-local Lee–Yang–Parr correlation functional [37], i.e., the unrestricted B3LYP/3-21G∼SDD approach. Moreover, the B3LYP/3-21G∼SDD optimized molecular geometries were further re-optimized using the standard 6-31G* basis set [38] for carbon atoms, i.e., the more advanced B3LYP/6-31G*∼SDD approach. The calculations are carried out for the quintet electronic state as the multiplicity produces the lowest energy at the B3LYP/6-31G*∼SDD computational level. The geometry optimizations at the B3LYP/6-31G*∼SDD level point out that the two isomers known from observations, cages C 2 v ( 9 ) and C s ( 6 ) , are indeed the lowest isomers in the potential energy (Table 1). Other isomers would have insignificant relative equilibrium populations owing to the enthalpy and entropy terms. In the optimized B3LYP/6-31G*∼SDD geometries, the harmonic vibrational analysis was carried out with the analytical force-constant matrix in order to generate the vibrational frequencies needed for the partition functions considered in the thermodynamic-stability description.
The SCF wavefunction stability [39,40,41] was checked throughout in order to prevent misleading unstable SCF solutions (that can happen rather frequently with metallofullerenes and could create a wrong stability picture if the wavefunction instability is not treated accordingly). The calculations were performed using the Gaussian 09 program [41] in parallel regime on computers operating with 8–24 processors (computational frequency up 3 GHz each and with the available operational memory up to 60 GB).
The relative equilibrium populations (expressed in the terms of mole fractions x i ) in our set of the two observed isomers can be expressed [42,43] with the help of their partition functions q i and the ground-state energies or enthalpies at absolute zero temperature Δ H 0 , i o (i.e., the relative potential energies corrected for the vibrational zero-point energies) by a formula:
x i = q i e x p [ Δ H 0 , i o / ( R T ) ] j = 1 2 q j e x p [ Δ H 0 , j o / ( R T ) ] .
The formula just presupposes the conditions of the inter-isomeric equilibrium (and thus—it cannot reflect a kinetic control if present in a reaction system). The rotational and vibrational partition functions are constructed [43] here with the conventional rigid-rotator and harmonic-oscillator (RRHO) approximation. No vibrational-frequency scaling is considered, as it is not significant [44,45] for the x i values at high temperatures. The temperature regions where metallofullerene electric-arc syntheses take place is not well known, however, the recent observations [46] suggest some arguments to expect it, for empty fullerenes, somewhere around 1300 K (though the region for metallofullerene syntheses could be somewhat lower owing to catalytic effects produced by the metals involved [47]).
For description of the metal motions in the cage, a modified [23,48] RRHO approach is however considered, respecting findings [49] that the encapsulated atoms can exhibit large amplitude vibrational motions, in particular at higher temperatures (interestingly, the motions can be restricted by cage derivatizations [50]). It can be expected that if the encapsulated atom is relatively free then, at high temperatures, its motions in different cages will yield about the same contribution to the partition functions. However, such comparable contributions should then cancel out in Equation (1). This approach is called the [23,48] free, fluctuating, or floating encapsulate model (FEM) and consists of two further simplifications. One is suppression of the three lowest vibrational frequencies (belonging to the metal motions in the cage), the other simplification takes symmetries of the cages as the highest (topologically) possible, which reflects averaging effects of the large-amplitude vibrational motions. For example, with the Dy@C82 isomer based on the IPR C 2 v ( 9 ) cage (Table 1), the C 2 v symmetry is considered within the FEM scheme though its static [51] symmetry (i.e., just resulting from the geometry optimizations) is only simple C 1 (Figure 1). Generally speaking, the FEM treatment is known to yield a better agreement [23,48] with observed data than the conventional RRHO approach. Hence, the FEM approach is also applied in this study.

3. Results and Discussion

Table 1 reports the Dy@C82 relative inter-isomeric energetics calculated at the B3LYP/6-31G*∼SDD level, i.e., either the difference in the potential energy, Δ E p o t , r e l , or the difference in the enthalpy at absolute zero temperature, Δ H 0 , r e l o . The two terms are related by the difference of zero-point vibrational energies Δ Z P E :
Δ H 0 , r e l o = Δ E p o t , r e l + Δ Z P E .
The Dy@ C 2 v ( 9 ) -C82 isomer is lower in the B3LYP/6-31G*∼SDD potential energy by 1.77 kcal/mol compared to the other stabilomer Dy@ C s ( 6 ) -C82 (Figure 1), or by 1.91 kcal in the enthalpy at 0 K scale. The two Dy@C82 isomers from Table 1 are considered in the following evaluation of their thermodynamic-equilibrium populations.
Table 2 presents selected calculated molecular characteristics of the two observed (and potential-energy-lowest) Dy@C82 isomers. The B3LYP/6-31G*∼SDD calculated closest contacts r D y C between the Dy atom and the cage are around 2.5 Å and hence rather similar to the values reported previously for other C82-based metallofullerenes [6,22,52]. The Mulliken atomic charge q D y on Dy evaluated at the B3LYP/3-21G∼SDD level is about 2.01 (in the units of the electron elementary charge), similar as found [6] for Er@C82. The charges on the cage carbons are both positive and negative (which is important for repulsions between the carbon atoms). The Dy@ C 2 v ( 9 ) -C82 isomer exhibits carbon charges from −0.201 to 0.010, Dy@ C s ( 6 ) -C82 from −0.205 to 0.012. The B3LYP/3-21G∼SDD calculated dipole moments of Dy@C82 are also similar to those found previously [6] for Er@C82. Such charge-distribution characteristics should help in analysis of the metallofullerene dopants for improvements of solar cells [5]. The lowest vibrational frequencies ω l o w presented in Table 2 are in agreement with the expected relatively free motion of metal atoms encapsulated in metallofullerenes. It should be mentioned that the charge distributions evaluated with the 3-21G∼SDD basis set (in contrast to, e.g., the 6-31G*∼SDD computational level) produce [32] for metallofullerenes a good agreement with the available observed charges [53]. Moreover, there are also general methodological reasons [54,55,56,57] why larger basis sets should not be used for evaluation of the charge distributions (larger basis sets can sometimes produce quite unphysical charge values [55]).
Figure 2 presents the key results of this study—temperature development of the relative equilibrium populations for the two studied Dy@C82 isomers in a wide temperature region. The relative populations are calculated with the FEM treatment based on the B3LYP/6-31G*∼SDD energy and entropy contributions. Clearly enough, at very low temperatures, the structure lower in the Δ H 0 , i o scale must prevail, i.e., the Dy@ C 2 v ( 9 ) -C82 species. However, the population of the other observed species Dy@ C s ( 6 ) -C82 increases rather fast so that both isomers have comparable concentrations at high temperatures around 1500 K (their precise equimolarity 50:50% is reached at 1570 K). At a supposed synthetic temperature [46,47] of 1000 K, the equilibrium populations are calculated as 58.0 and 42.0% for the Dy@ C 2 v ( 9 ) -C82 and Dy@ C s ( 6 ) -C82 isomer, respectively. The FEM B3LYP/6-31G*∼SDD calculations thus agree with the observations [8,9,10,11,12,13,14,15,16] of the two cage symmetries, C 2 v ( 9 ) (major isomer) and C s ( 6 ) (minor isomer). Interestingly, when the entropic contributions are completely neglected in the evaluation of the relative populations (i.e., if just the simple Boltzmann factors [42,43] are considered), the relative isomeric population at the same temperature of 1000 K is equal to 71.0 and 29.0% for the Dy@ C 2 v ( 9 ) -C82 and Dy@ C s ( 6 ) -C82 isomer, respectively (while in the conventional RRHO approximation the isomeric populations differ even more from the recommended FEM values—the RRHO terms are 82.0 and 18%).
It should be mentioned that although Equation (1) deals with just the two isomers we are interested in, the ratio itself of their relative populations would not be changed if they would be placed into the equilibrium mixture with more isomers (i.e., the two relative poulations would be changed, however, not their ratio). If more isomers would be considered, there would be a different denominator in Equation (1). Still, when a ratio of the two isomers is evaluated, the denominator would be just cancelled out. In particular, the two isomers even in presence of some other isomers would again exhibit their equimolarity at the temperature of 1570 K (as found for the two isomers alone—Figure 2)—hence, this equimolarity temperature represents a kind of transferable parameter.
The B3LYP/6-31G*∼SDD vibrational frequencies serve here primarily in the partitions functions in Equation (1), however, they can also be used for simulation of the IR vibrational spectra (Figure 3). The IR spectra could be used for identification of the isomers.
Some similarities are to be expected with the results previously obtained [18,30,58,59,60,61,62,63] in calculations of other C82-based metallofullerenes with a comparable metal-to-cage charge transfer. Such similarities can be related to the aspect that metallofullerenes are not formed by a covalent but by ionic bond [64,65,66]. There are however some general problems with comparison of calculated and observed data. The observed isomeric populations can depend on applied metal form [67]. The feature can be related to catalytic and kinetic factors [47,68,69]. Moreover, it is generally difficult to check if the supposed inter-isomeric thermodynamic equilibrium is really achieved in observations. Yet, another experimental issue is possible with different solubility [70,71,72] of various endohedral isomers in the applied extraction solvent.

4. Conclusions

The calculations of the relative equilibrium populations for the two observed IPR isomers of Dy@C82 in the temperature region used [46] in metallofullerene syntheses, employing the Gibbs energy (which is based on quantum-chemical molecular parameters), agree with available observed data. Still, it can be interesting from the methodological point of view to check the results at yet higher quantum-chemical levels as well as to improve Equation (1) with a higher description of the partition functions—when technically possible. Overall, the results encourage applications of the Gibbs-energy approach to the equilibrium populations of even more complex nanocarbon systems (e.g., [73,74,75,76,77,78,79,80,81,82,83,84,85]). In particular, further calculations are in progress in order to obtain a more detailed understanding of the functions of fullerene endohedrals as dopants for higher performance and stability of solar cells [6,86,87,88]. In addition to Dy@C82 [1,2], also Er@C82 [5,6] and Gd@C82 [61,87,88,89] are studied, especially their molecular and electronic structures as well as their energetics, thermodynamics, and also hydrophobicity [5,90,91,92,93,94,95]. The theoretical treatments should further be used in the formulation of molecular rules for rational screening of fullerene endohedrals for their applications in photovoltaics.

Author Contributions

Conceptualization, Z.S., T.A. and X.L.; methodology, Z.S. and L.A.; hardware and software, Z.S., F.U. and L.A.; models validation, Z.S., T.A., X.L. and L.A.; analysis and interpretation, Z.S., F.U., T.A. and X.L.; writing—original draft preparation, Z.S. and F.U. All authors have read and agreed to the published version of the manuscript.

Funding

The reported research has been supported by the National Natural Science Foundation of China (21925104 and 92261204), the Hubei Provincial Natural Science Foundation of China (No. 2021CFA020), the International Cooperation Key Project of Science and Technology Department of Shaanxi, the Charles University Centre of Advanced Materials/CUCAM (CZ.02.1.01/0.0/0.0/15 003/0000417), and the MetaCentrum ((LM2010005) and CERIT-SC (CZ.1.05/3.2.00/08.0144) computing facilities. A very initial phase of the research line was supported by the Alexander von Humboldt-Stiftung and the Max-Planck-Institut für Chemie (Otto-Hahn-Institut), too.

Data Availability Statement

The data presented in this study are available in article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, S.; Fan, L.; Yang, S. Preparation, characterization, and photoelectrochemistry of Langmuir-Blodgett films of the endohedral metallofullerene Dy@C82 mixed with metallophthalocyanines. J. Phys. Chem. B 2003, 107, 8403–8411. [Google Scholar] [CrossRef]
  2. Yang, S.; Fan, L.; Yang, S. Langmuir–Blodgett films of poly(3-hexylthiophene) doped with the endohedral metallofullerene Dy@C82: Preparation, characterization, and application in photoelectrochemical cells. J. Phys. Chem. B 2004, 108, 4394–4404. [Google Scholar] [CrossRef]
  3. Yang, S.; Fan, L.; Yang, Y. Significantly enhanced photocurrent efficiency of a poly(3-hexylthiophene) photoelectrochemical device by doping with the endohedral metallofullerene Dy@C82. Chem. Phys. Lett. 2004, 388, 253–258. [Google Scholar] [CrossRef]
  4. Katz, E.A. Fullerene thin films as photovoltaic material. In Nanostructured Materials for Solar Energy Conversion; Soga, T., Ed.; Elsevier: Amsterdam, The Netherlands, 2006; pp. 361–443. [Google Scholar]
  5. Ye, X.; Yu, P.; Shen, W.; Hu, S.; Akasaka, T.; Lu, X. Er@C82 as a bifunctional additive to the spiro-OMeTAD hole transport layer for improving performance and stability of perovskite solar cells. Sol. RRL 2021, 5, 2100463. [Google Scholar] [CrossRef]
  6. Slanina, Z.; Uhlík, F.; Hu, S.; Akasaka, T.; Lu, X.; Adamowicz, L. Calculated isomeric populations of Er@C82. Fullerenes Nanotub. Carbon Nanostruct. 2024, 32, 986–991. [Google Scholar] [CrossRef]
  7. Mumyatov, A.V.; Troshin, P.A. A review on fullerene derivatives with reduced electron affinity as acceptor materials for organic solar cells. Energies 2023, 16, 1924. [Google Scholar] [CrossRef]
  8. Iida, S.; Kubozono, Y.; Slovokhotov, Y.; Takabayashi, Y.; Kanbara, T.; Fukunaga, T.; Fujiki, S.; Emura, S.; Kashino, S. Structure and electronic properties of Dy@C82 studied by UV–VIS absorption, X-ray powder diffraction and XAFS. Chem. Phys. Lett. 2001, 338, 21–28. [Google Scholar]
  9. Takabayashi, Y.; Haruyama, Y.; Rikiishi, Y.; Hosokawa, T.; Shibata, K.; Kubozono, Y. Preferred location of the Dy ion in the minor isomer of Dy@C82 determined by Dy LIII-edge EXAFS. Chem. Phys. Lett. 2004, 388, 23–26. [Google Scholar]
  10. Xenogiannopoulou, E.; Couris, S.; Koudoumas, E.; Tagmatarchis, N.; Inoue, T.; Shinohara, H. Nonlinear optical response of some isomerically pure higher fullerenes and their corresponding endohedral metallofullerene derivatives: C82–C2v, Dy@C82(I), Dy2@C82(I), C92–C2 and Er2@C92(IV). Chem. Phys. Lett. 2004, 394, 14. [Google Scholar]
  11. Li, X.; Fan, L.; Liu, D.; Sung, H.H.Y.; Williams, I.D.; Yang, S.; Tan, K.; Lu, X. Synthesis of a Dy@C82 derivative bearing a single phosphorus substituent via a zwitterion approach. J. Am. Chem. Soc. 2007, 129, 10636–10637. [Google Scholar] [CrossRef]
  12. Kitaura, R.; Okimoto, H.; Shinohara, H.; Nakamura, T.; Osawa, H. Magnetism of the endohedral metallofullerenes M@C82 (M = Gd, Dy) and the corresponding nanoscale peapods: Synchrotron soft x-ray magnetic circular dichroism and density-functional theory calculations. Phys. Rev. B 2007, 76, 172409. [Google Scholar] [CrossRef]
  13. Rodríguez-Fortea, A.; Balch, A.L.; Poblet, J.M. Endohedral metallofullerenes: A unique host-guest association. Chem. Soc. Rev. 2011, 40, 3551–3563. [Google Scholar] [CrossRef]
  14. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  15. Kareev, I.E.; Nekrasov, V.M.; Dutlov, A.E.; Bubnov, V.P.; Martynenko, V.M.; Laukhina, E.E.; Veciana, J.; Rovira, C. Determination of molar extinction coefficients for endohedral metallofullerene Dy@C82(C2v). Russ. Chem. Bull. (Int. Ed.) 2016, 65, 2421–2424. [Google Scholar] [CrossRef]
  16. Yang, W.; Barbosa, M.F.S.; Alfonsov, A.; Rosenkranz, M.; Israel, N.; Büchner, B.; Avdoshenko, S.M.; Liu, F.; Popov, A.A. Thirty years of hide-and-seek: Capturing abundant but elusive MIII@C3v(8)-C82 isomer, and the study of magnetic anisotropy induced in Dy3+ ion by the fullerene π-ligand. J. Am. Chem. Soc. 2024, 146, 25328–25342. [Google Scholar] [CrossRef] [PubMed]
  17. Slanina, Z.; Uhlík, F.; Feng, L.; Adamowicz, L. Calculated relative populations of Sm@C82 isomers. Fullerenes Nanotub. Carbon Nanostruct. 2018, 26, 233–238. [Google Scholar] [CrossRef]
  18. Slanina, Z.; Lee, S.-L.; Kobayashi, K.; Nagase, S. AM1 computed thermal effects within the nine isolated-pentagon-rule isomers of C82. J. Mol. Struct. (Theochem) 1995, 339, 89–93. [Google Scholar] [CrossRef]
  19. Nishibori, E.; Takata, M.; Sakata, M.; Inakuma, M.; Shinohara, H. Determination of the cage structure of Sc@C82 by synchrotron powder diffraction. Chem. Phys. Lett. 1998, 298, 79–84. [Google Scholar] [CrossRef]
  20. Slanina, Z.; Kobayashi, K.; Nagase, S. Ca@C82 isomers: Computed temperature dependency of relative concentrations. J. Chem. Phys. 2004, 120, 3397–3400. [Google Scholar] [CrossRef] [PubMed]
  21. Slanina, Z.; Kobayashi, K.; Nagase, S. Computed temperature development of the relative stabilities of La@C82 isomers. Chem. Phys. Lett. 2004, 388, 74–78. [Google Scholar] [CrossRef]
  22. Suzuki, M.; Slanina, Z.; Mizorogi, N.; Lu, X.; Nagase, S.; Olmstead, M.M.; Balch, A.L.; Akasaka, T. Single-crystal X-ray diffraction study of three Yb@C82 isomers cocrystallized with Ni-II(octaethylporphyrin). J. Am. Chem. Soc. 2012, 134, 18772–18778. [Google Scholar] [CrossRef]
  23. Hu, Z.; Hao, Y.; Slanina, Z.; Gu, Z.; Shi, Z.; Uhlík, F.; Zhao, Y.; Feng, L. Popular C82 fullerene cage encapsulating a divalent metal ion Sm2+: Structure and electrochemistry. Inorg. Chem. 2015, 54, 2103–2108. [Google Scholar] [CrossRef]
  24. Slanina, Z.; Lee, S.-L.; Adamowicz, L. C80, C86, C88: Semiempirical and ab initio SCF calculations. Int. J. Quantum. Chem. 1997, 63, 529–535. [Google Scholar] [CrossRef]
  25. Slanina, Z.; Uhlík, F. Temperature dependence of the Gibbs energy ordering of isomers of Cl2O2. J. Phys. Chem. 1991, 95, 5432–5434. [Google Scholar] [CrossRef]
  26. Slanina, Z.; Zhao, X.; Lee, S.-L.; Ōsawa, E. C90 Temperature effects on relative stabilities of the IPR isomers. Chem. Phys. 1997, 219, 193–200. [Google Scholar] [CrossRef]
  27. Uhlík, F.; Slanina, Z.; Ōsawa, E. C78 IPR fullerenes: Computed B3LYP/6-31G*//HF/3-21G temperature-dependent relative concentrations. Eur. Phys. J. D 2001, 16, 349–352. [Google Scholar] [CrossRef]
  28. Slanina, Z.; Zhao, X.; Uhlík, F.; Lee, S.-L.; Adamowicz, L. Computing enthalpy-entropy interplay for isomeric fullerenes. Int. J. Quantum Chem. 2004, 99, 640–653. [Google Scholar] [CrossRef]
  29. Slanina, Z.; Lee, S.-L.; Adamowicz, L.; Uhlík, F.; Nagase, S. Computed structure and energetics of La@C60. Int. J. Quantum Chem. 2005, 104, 272–277. [Google Scholar] [CrossRef]
  30. Slanina, Z.; Lee, S.-L.; Uhlík, F.; Adamowicz, L.; Nagase, S. Computing relative stabilities of metallofullerenes by Gibbs energy treatments. Theor. Chem. Acc. 2007, 117, 315–322. [Google Scholar] [CrossRef]
  31. Wang, Y.; Morales-Martínez, R.; Zhang, X.; Yang, W.; Wang, Y.; Rodríguez-Fortea, A.; Poblet, J.M.; Feng, L.; Wang, S.; Chen, N. Unique four-electron metal-to-cage charge transfer of Th to a C82 fullerene cage: Complete structural characterization of Th@C3v(8)-C82. J. Am. Chem. Soc. 2017, 139, 5110–5116. [Google Scholar] [CrossRef]
  32. Slanina, Z.; Uhlík, F.; Nagase, S.; Akasaka, T.; Adamowicz, L.; Lu, X. Eu@C72: Computed comparable populations of two non-IPR isomers. Molecules 2017, 22, 1053. [Google Scholar]
  33. Zhao, Y.; Li, M.; Zhao, R.; Zhao, P.; Yuan, K.; Li, Q.; Zhao, X. Unmasking the optimal isomers of Ti2C84: Ti2C2@C82 Instead of Ti2C84. J. Phys. Chem. C 2018, 122, 13148–13155. [Google Scholar] [CrossRef]
  34. Binkley, J.S.; Pople, J.A.; Hehre, W.J. Self-consistent molecular orbital methods. 21. Small split-valence basis sets for first-row elements. J. Am. Chem. Soc. 1980, 102, 939–947. [Google Scholar] [CrossRef]
  35. Cao, X.Y.; Dolg, M. Segmented contraction scheme for small-core lanthanide pseudopotential basis sets. J. Mol. Struct. (Theochem) 2002, 581, 139–147. [Google Scholar] [CrossRef]
  36. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  37. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  38. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular-orbital studies of organic-molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  39. Schlegel, H.B.; McDouall, J.J.W. Do you have SCF stability and convergence problems? In Computational Advances in Organic Chemistry; Ögretir, C., Csizmadia, I.G., Eds.; Kluwer: Dordrecht, The Netherlands, 1991; pp. 167–185. [Google Scholar]
  40. Slanina, Z.; Uhlík, F.; Adamowicz, L. Computations of model narrow nanotubes closed by fragments of smaller fullerenes and quasi-fullerenes. J. Mol. Graph. Mod. 2003, 21, 517–522. [Google Scholar] [CrossRef] [PubMed]
  41. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Rev. C.01; Gaussian Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  42. Slanina, Z. Equilibrium isomeric mixtures: Potential energy hypersurfaces as originators of the description of the overall thermodynamics and kinetics. Int. Rev. Phys. Chem. 1987, 6, 251–267. [Google Scholar] [CrossRef]
  43. Slanina, Z. A Program for determination of composition and thermodynamics of the ideal gas-phase equilibrium isomeric mixtures. Comput. Chem. 1989, 13, 305–311. [Google Scholar] [CrossRef]
  44. Slanina, Z.; Uhlík, F.; Zerner, M.C. C 5 H 3 + isomeric structures: Relative stabilities at high temperatures. Rev. Roum. Chim. 1991, 36, 965–974. [Google Scholar]
  45. Slanina, Z.; Adamowicz, L. On relative stabilities of dodecahedron-shaped and bowl-shaped structures of C20. Thermochim. Acta 1992, 205, 299–306. [Google Scholar] [CrossRef]
  46. Cross, R.J.; Saunders, M. Transmutation of fullerenes. J. Am. Chem. Soc. 2005, 127, 3044–3047. [Google Scholar] [CrossRef] [PubMed]
  47. Slanina, Z.; Zhao, X.; Uhlík, F.; Ozawa, M.; Osawa, E. Computational modelling of the metal and other elemental catalysis in the Stone-Wales fullerene rearrangements. J. Organomet. Chem. 2000, 599, 57–61. [Google Scholar] [CrossRef]
  48. Slanina, Z.; Adamowicz, L.; Kobayashi, K.; Nagase, S. Gibbs energy-based treatment of metallofullerenes: Ca@C72, Ca@C74, Ca@C82, and La@C82. Mol. Simul. 2005, 31, 71–77. [Google Scholar] [CrossRef]
  49. Akasaka, T.; Nagase, S.; Kobayashi, K.; Walchli, M.; Yamamoto, K.; Funasaka, H.; Kako, M.; Hoshino, T.; Erata, T. 13C and 139La NMR studies of La2@C80: First evidence for circular motion of metal atoms in endohedral dimetallofullerenes. Angew. Chem. Int. Ed. 1997, 36, 1643–1645. [Google Scholar] [CrossRef]
  50. Kobayashi, K.; Nagase, S.; Maeda, Y.; Wakahara, T.; Akasaka, T. La2@C80: Is the circular motion of two La atoms controllable by exohedral addition? Chem. Phys. Lett. 2003, 374, 562–566. [Google Scholar] [CrossRef]
  51. Slanina, Z. Contemporary Theory of Chemical Isomerism; Academia: Prague, Czech Republic; D. Reidel Publishing Company: Dordrecht, The Netherlands, 1986; p. 21. [Google Scholar]
  52. Slanina, Z.; Uhlík, F.; Bao, L.; Akasaka, T.; Lu, X.; Adamowicz, L. Calculated relative populations for the Eu@C82 isomers. Chem. Phys. Lett. 2019, 726, 29–33. [Google Scholar] [CrossRef]
  53. Takata, M.; Nishibori, E.; Sakata, M.; Shinohara, H. Charge density level structures of endohedral metallofullerenes determined by synchrotron radiation powder method. New Diam. Front. Carb. Technol. 2002, 12, 271–286. [Google Scholar]
  54. Hehre, W.J. A Guide to Molecular Mechanics and Quantum Chemical Calculations; Wavefunction: Irvine, CA, USA, 2003; p. 435. [Google Scholar]
  55. Jensen, F. Introduction to Computational Chemistry; Wiley: Chichester, UK, 2017; p. 319. [Google Scholar]
  56. Campanera, J.M.; Bo, C.; Poblet, J.M. General rule for the stabilization of fullerene cages encapsulating trimetallic nitride templates. Angew. Chem. Int. Ed. 2005, 44, 7230–7233. [Google Scholar] [CrossRef]
  57. Slanina, Z.; Uhlík, F.; Pan, C.; Akasaka, T.; Lu, X.; Adamowicz, L. Computed stabilization for a giant fullerene endohedral: Y2C2@C1(1660)-C108. Chem. Phys. Lett. 2018, 710, 147–149. [Google Scholar] [CrossRef]
  58. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Nagase, S. Structural and bonding features of Z@C82 (Z = Al, Sc, Y, La) endohedrals. J. Comput. Meth. Sci. Eng. 2010, 10, 569–574. [Google Scholar] [CrossRef]
  59. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Suzuki, M.; Lu, X.; Mizorogi, N.; Nagase, S.; Akasaka, T. Calculated temperature development of the relative stabilities of Yb@C82 isomers. Fuller. Nanotub. Car. Nanostruct. 2014, 22, 147–154. [Google Scholar] [CrossRef]
  60. Slanina, Z.; Uhlík, F.; Shen, W.; Akasaka, T.; Lu, X.; Adamowicz, L. Calculations of the relative populations of Lu@C82 isomers. Fullerenes Nanotub. Carbon Nanostruct. 2019, 27, 710–714. [Google Scholar] [CrossRef]
  61. Slanina, Z.; Uhlík, F.; Akasaka, T.; Lu, X.; Adamowicz, L. Calculated relative thermodynamic stabilities of the Gd@C82 isomers. ECS J. Solid State Sci. Technol. 2021, 10, 071013. [Google Scholar] [CrossRef]
  62. Meng, Q.Y.; Morales-Martínez, R.; Zhuang, J.X.; Yao, Y.R.; Wang, Y.F.; Feng, L.; Poblet, J.M.; Rodríguez-Fortea, A.; Chen, N. Synthesis and characterization of two isomers of Th@C82: Th@C2v(9)-C82 and Th@C2(5)-C82. Inorg. Chem. 2021, 60, 11496–11502. [Google Scholar] [CrossRef]
  63. Slanina, Z.; Uhlík, F.; Feng, L.; Adamowicz, L. Ho@C82 metallofullerene: Calculated isomeric composition. ECS J. Solid State Sci. Technol. 2022, 11, 053018. [Google Scholar] [CrossRef]
  64. Andreoni, W.; Curioni, A. Freedom and constraints of a metal atom encapsulated in fullerene cages. Phys. Rev. Lett. 1996, 77, 834–837. [Google Scholar] [CrossRef]
  65. Popov, A.A.; Dunsch, L. Bonding in endohedral metallofullerenes as studied by quantum theory of atoms in molecules. Chem. Eur. J. 2009, 15, 9707–9729. [Google Scholar] [CrossRef] [PubMed]
  66. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Adamowicz, L.; Akasaka, T.; Nagase, S. Computed stabilities in metallofullerene series: Al@C82, Sc@C82, Y@C82, and La@C82. Int. J. Quant. Chem. 2011, 111, 2712–2718. [Google Scholar] [CrossRef]
  67. Yang, H.; Yu, M.; Jin, H.; Liu, Z.; Yao, M.; Liu, B.; Olmstead, M.M.; Balch, A.L. Isolation of three isomers of Sm@C84 and X-ray crystallographic characterization of Sm@D3d(19)-84 and Sm@C2(13)-C84. J. Am. Chem. Soc. 2012, 134, 5331–5338. [Google Scholar] [CrossRef]
  68. Hao, Y.; Feng, L.; Xu, W.; Gu, Z.; Hu, Z.; Shi, Z.; Slanina, Z.; Uhlík, F. Sm@C2v(19138)-C76: A non-IPR cage stabilized by a divalent metal ion. Inorg. Chem. 2015, 54, 4243–4248. [Google Scholar] [CrossRef] [PubMed]
  69. Hao, Y.; Tang, Q.; Li, X.; Zhang, M.; Wan, Y.; Feng, L.; Chen, N.; Slanina, Z.; Adamowicz, L.; Uhlík, F. Isomeric Sc2O@C78 related by a single-step Stone–Wales transformation: Key links in an unprecedented fullerene formation pathway. Inorg. Chem. 2016, 55, 11354–11361. [Google Scholar] [CrossRef] [PubMed]
  70. Jehlička, J.; Svatoš, A.; Frank, O.; Uhlík, F. Evidence for fullerenes in solid bitumen from pillow lavas of proterozoic age from Mítov (Bohemian Massif, Czech Republic). Geochem. Cosmochem. Acta 2003, 67, 1495–1506. [Google Scholar] [CrossRef]
  71. Lian, Y.; Shi, Z.; Zhou, X.; Gu, Z. Different extraction behaviors between divalent and trivalent endohedral metallofullerenes. Chem. Mater. 2004, 16, 1704–1714. [Google Scholar] [CrossRef]
  72. Maeda, Y.; Tsuchiya, T.; Kikuchi, T.; Nikawa, H.; Yang, T.; Zhao, X.; Slanina, Z.; Suzuki, M.; Yamada, M.; Lian, Y.; et al. Effective derivatization and extraction of insoluble missing lanthanum metallofullerenes La@C2n (n = 36–38) with iodobenzene. Carbon 2016, 98, 67–73. [Google Scholar] [CrossRef]
  73. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Adamowicz, L.; Nagase, S. Computations of endohedral fullerenes: The Gibbs energy treatment. J. Comput. Meth. Sci. Engn. 2006, 6, 243–250. [Google Scholar] [CrossRef]
  74. Slanina, Z.; Uhlík, F.; Akasaka, T.; Lu, X.; Adamowicz, L. Theoretical studies of non-metal endohedral fullerenes. Nanomaterials 2025, 15, 1287. [Google Scholar] [CrossRef]
  75. Fang, Y.; Bi, C.; Wang, D.; Huang, J. The functions of fullerenes in hybrid perovskite solar cells. ACS Energy Lett. 2017, 2, 782–794. [Google Scholar] [CrossRef]
  76. Okazaki, T.; Shimada, T.; Suenaga, K.; Ohno, Y.; Mizutani, T.; Lee, J.; Kuk, Y.; Shinohara, H. Electronic properties of Gd@C82 metallofullerene peapods: (Gd@C82)n@SWNTs. Appl. Phys. A 2003, 76, 475–478. [Google Scholar] [CrossRef]
  77. Zhang, K.K.; Wang, C.; Zhang, M.H.; Bai, Z.B.; Xie, F.F.; Tan, Y.Z.; Guo, Y.L.; Hu, K.J.; Cao, L.; Zhang, S.; et al. A Gd@C82 single-molecule electret. Nat. Nanotech. 2020, 15, 1019–1024. [Google Scholar] [CrossRef]
  78. Wu, B.-S.; An, M.-W.; Chen, J.-M.; Xing, Z.; Chen, Z.-C.; Deng, L.-L.; Tian, H.-R.; Yun, D.-Q.; Xie, S.-Y.; Zheng, L.-S. Radiation-processed perovskite solar cells with fullerene-enhanced performance and stability. Cell Rep. Phys. Sci. 2021, 2, 100646. [Google Scholar] [CrossRef]
  79. Leenaerts, O.; Partoens, B.; Peeters, F.M. Water on graphene: Hydrophobicity and dipole moment using density functional theory. Phys. Rev. B 2009, 79, 235440. [Google Scholar] [CrossRef]
  80. Rašović, I. Water-soluble fullerenes for medical applications. Mater. Sci. Tech. 2017, 33, 777–794. [Google Scholar] [CrossRef]
  81. Bologna, F.; Mattioli, E.J.; Bottoni, A.; Zerbetto, F.; Calvaresi, M. Interactions between endohedral metallofullerenes and proteins: The Gd@C60–lysozyme model. ACS Omega 2018, 3, 13782–13789. [Google Scholar] [CrossRef]
  82. Meng, H.; Zhao, C.; Nie, M.; Wang, C.; Wang, T. Changing the hydrophobic MOF pores through encapsulating fullerene C60 and metallofullerene Sc3C2@C80. J. Phys. Chem. C 2019, 123, 6265–6269. [Google Scholar] [CrossRef]
  83. Grebowski, J.; Litwinienko, G. Metallofullerenols in biomedical applications. Eur. J. Med. Chem. 2022, 238, 114481. [Google Scholar] [CrossRef] [PubMed]
  84. Matnazarova, S.; Matyakubov, N.; Khalilov, U.; Yusupov, M. Influence of encapsulated transition metal atoms on the hydrophilic properties of C84 fullerene: A computational study. Chem. Phys. Lett. 2025, 869, 142067. [Google Scholar] [CrossRef]
  85. Gueorguiev, G.K.; Stafström, S.; Hultman, L. Nano-wire formation by self-assembly of silicon-metal cage-like molecules. Chem. Phys. Lett. 2008, 458, 170–174. [Google Scholar] [CrossRef]
  86. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Akasaka, T.; Nagase, S. Stability computations for fullerenes and metallofullerenes. In Handbook of Carbon Nano Materials; D’Souza, F., Kadish, K.M., Eds.; Materials and Fundamental Applications; World Scientific Publishing Co.: Singapore, 2012; Volume 4, pp. 381–429. [Google Scholar]
  87. An, D.-Y.; Su, J.-G.; Li, C.-H.; Li, J.-Y. Computational studies on the interactions of nanomaterials with proteins and their impacts. Chin. Phys. B 2015, 24, 120504. [Google Scholar] [CrossRef]
  88. Basiuk, V.A.; Tahuilan-Anguiano, D.E. Complexation of free-base and 3d transition metal(II) phthalocyanines with endohedral fullerene Sc3N@C80. Chem. Phys. Lett. 2019, 722, 146–152. [Google Scholar] [CrossRef]
  89. Tahuilan-Anguiano, D.E.; Basiuk, V.A. Complexation of free-base and 3d transition metal(II) phthalocyanines with endohedral fullerenes H@C60, H2@C60 and He@C60: The effect of encapsulated species. Diam. Relat. Mater. 2021, 118, 108510. [Google Scholar] [CrossRef]
  90. Uhlík, F.; Slanina, Z.; Bao, L.; Akasaka, T.; Lu, X.; Adamowicz, L. Eu@C88 isomers: Calculated relative populations. ECS J. Solid State Sci. Technol. 2022, 11, 101008. [Google Scholar] [CrossRef]
  91. Li, M.; Zhao, R.; Dang, J.; Zhao, X. Theoretical study on the stabilities, electronic structures, and reaction and formation mechanisms of fullerenes and endohedral metallofullerenes. Coord. Chem. Rev. 2022, 471, 214762. [Google Scholar] [CrossRef]
  92. Slanina, Z.; Uhlík, F.; Adamowicz, L. Theoretical predictions of fullerene stabilities. In Handbook of Fullerene Science and Technology; Lu, X., Akasaka, T., Slanina, Z., Eds.; Springer: Singapore, 2022; pp. 111–179. [Google Scholar]
  93. Menon, A.; Kaur, R.; Guldi, D.M. Merging Carbon Nanostructures with Porphyrins. In Handbook of Fullerene Science and Technology; Lu, X., Akasaka, T., Slanina, Z., Eds.; Springer: Singapore, 2022; pp. 219–264. [Google Scholar]
  94. Basiuk, V.A.; Wu, Y.F.; Prezhdo, O.V.; Basiuk, E.V. Lanthanide atoms induce strong graphene sheet distortion when adsorbed on Stone-Wales defects. J. Phys. Chem. Lett. 2024, 15, 9706–9713. [Google Scholar] [CrossRef]
  95. Gra̧dzka, E. Electrocatalytic properties of fullerene-based materials. In NanoCarbon: A Wonder Material for Energy Applications; Gupta, R.K., Ed.; Springer: Singapore, 2024; Volume 1, pp. 199–218. [Google Scholar]
Figure 1. The B3LYP/6-31G*∼SDD optimized structures of the two energy-lowest Dy@C82 isomers, Dy@ C 2 v ( 9 ) -C82 left, Dy@ C s ( 6 ) -C82 right (the shortest Dy-C contact is indicated by a link).
Figure 1. The B3LYP/6-31G*∼SDD optimized structures of the two energy-lowest Dy@C82 isomers, Dy@ C 2 v ( 9 ) -C82 left, Dy@ C s ( 6 ) -C82 right (the shortest Dy-C contact is indicated by a link).
Micro 05 00049 g001
Figure 2. The equilibrium relative populations of the Dy@C82 isomers based on the FEM treatment at the B3LYP/6-31G*∼SDD level.
Figure 2. The equilibrium relative populations of the Dy@C82 isomers based on the FEM treatment at the B3LYP/6-31G*∼SDD level.
Micro 05 00049 g002
Figure 3. B3LYP/6-31G*∼SDD calculated IR spectrum of Dy@ C s ( 6 ) -C82 (top) and Dy@ C 2 v ( 9 ) -C82 (bottom).
Figure 3. B3LYP/6-31G*∼SDD calculated IR spectrum of Dy@ C s ( 6 ) -C82 (top) and Dy@ C 2 v ( 9 ) -C82 (bottom).
Micro 05 00049 g003
Table 1. B3LYP/6-31G*∼SDD relative potential energies Δ E p o t , r e l and enthalpies at 0 K Δ H 0 , r e l o of the two energy-lowest Dy@C82 isomers.
Table 1. B3LYP/6-31G*∼SDD relative potential energies Δ E p o t , r e l and enthalpies at 0 K Δ H 0 , r e l o of the two energy-lowest Dy@C82 isomers.
Species Δ E pot , rel /kcal·mol−1 Δ H 0 , rel o /kcal·mol−1
C s ( 6 )  a1.771.91
C 2 v ( 9 )  a0.00.0
a See Figure 1.
Table 2. The selected characteristics of the two energy-lowest Dy@C82 isomers—the closest Dy-C contact a r D y C , the Mulliken charge b on Dy q D y , the dipole moment b p, the lowest vibrational frequency a ω l o w .
Table 2. The selected characteristics of the two energy-lowest Dy@C82 isomers—the closest Dy-C contact a r D y C , the Mulliken charge b on Dy q D y , the dipole moment b p, the lowest vibrational frequency a ω l o w .
Species r Dy C q Dy p/Debye ω low /cm−1
C s ( 6 ) c2.4802.0180.80127.5
C 2 v ( 9 ) c2.5132.0111.34518.2
a B3LYP/6-31G*∼SDD terms. b B3LYP/3-21G∼SDD terms. c See Figure 1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Slanina, Z.; Uhlík, F.; Akasaka, T.; Lu, X.; Adamowicz, L. A Note on Computational Characterization of Dy@C82: Dopant for Solar Cells. Micro 2025, 5, 49. https://doi.org/10.3390/micro5040049

AMA Style

Slanina Z, Uhlík F, Akasaka T, Lu X, Adamowicz L. A Note on Computational Characterization of Dy@C82: Dopant for Solar Cells. Micro. 2025; 5(4):49. https://doi.org/10.3390/micro5040049

Chicago/Turabian Style

Slanina, Zdeněk, Filip Uhlík, Takeshi Akasaka, Xing Lu, and Ludwik Adamowicz. 2025. "A Note on Computational Characterization of Dy@C82: Dopant for Solar Cells" Micro 5, no. 4: 49. https://doi.org/10.3390/micro5040049

APA Style

Slanina, Z., Uhlík, F., Akasaka, T., Lu, X., & Adamowicz, L. (2025). A Note on Computational Characterization of Dy@C82: Dopant for Solar Cells. Micro, 5(4), 49. https://doi.org/10.3390/micro5040049

Article Metrics

Back to TopTop