Next Article in Journal
Accurate and Scalable Quantum Hydrodynamic Simulations of Plasmonic Nanostructures Within OFDFT
Previous Article in Journal
Nanostructured Higher Manganese Silicide Thermoelectrics Developed by Mechanical Alloying Using High-Purity and Recycled Silicon
Previous Article in Special Issue
Optimization of Solid Lipid Nanoparticle Formulation Using Design of Experiments, PART I: Strategic Tool for the Determination of Critical Parameters Regarding Formulation and Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Theoretical Studies of Non-Metal Endohedral Fullerenes

1
Department of Chemistry and Biochemistry, University of Arizona, Tucson, AZ 85721, 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, Albertov 6, 128 43 Praha 2, Czech Republic
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(16), 1287; https://doi.org/10.3390/nano15161287
Submission received: 25 March 2025 / Revised: 26 April 2025 / Accepted: 30 April 2025 / Published: 21 August 2025
(This article belongs to the Special Issue Modeling, Simulation and Optimization of Nanomaterials)

Abstract

This article presents computational studies of non-metal fullerene endohedrals, which are useful for understanding and interpreting experimental results. The encapsulated non-metal species are simple molecules like H2, N2, CO, HF, NH3, H2O2, H2O, and their aggregates. Predictions of thermodynamic stability and reaction populations are reviewed, based on quantum-chemical and statistical–thermodynamic treatments. As fullerene syntheses are performed at high temperatures, some of the calculations are based on both the encapsulation potential energy and the encapsulation Gibbs energy changes.

1. Introduction

Recently, endohedral encapsulation in fullerenes has been extended from metals to non-metal species (Table 1 presents illustrative examples). Metallofullerenes are produced by internal charge transfer. However, the charge transfer is not important for encapsulations of non-metal species—their formation is related [1] to weak interactions. N2@C60 and N2@C70 are examples of such non-metal endohedral fullerenes, produced [2] by means of high-temperature and high-pressure treatment. In fact, among two thousand C60 molecules, approximately one incorporates N2 [2]. Endohedrals with a nitrogen molecule can survive even several hours of heating at higher temperatures. The N2@C60 species is also observed [3] in the chromatographic analysis of nitrogen implantation in C60, otherwise primarily yielding N@C60 [4,5,6,7,8,9]. Endohedral fullerenes containing noble atoms [10,11,12,13,14] are also produced [10] by means of high-temperature and high-pressure techniques with a catalyst [13]. A more recent synthetic method for the encapsulation of non-metallic species (for example, hydrogen molecules [15] or water molecules [16]) put the molecule first in an open carbon cage and then its window was closed [17,18] by synthetic means. For example, complex synthesis produced [19] (H2O)2@C70. Carbon monoxide [20,21] or H2O2 [22,23] were also encapsulated in open-cage C60 derivatives. Obviously, such new fullerene encapsulations have also been calculated [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45]—such theoretical treatments are surveyed in this report and illustrated with topical systems.

2. H2 Encapsulation in C60

Originally, non-metallic fullerene encapsulates were calculated primarily [30] with the traditional B3LYP functional and later on using other more recent DFT functionals [47,48,49,50] (the 6–31G*, 6–31G**, and other standard atomic basis sets). The calculations can be augmented with the evaluation of a correction known as a basis set superposition error (BSSE). The BSSE term is estimated using the counterpoise technique introduced by Boys and Bernardi [51]. BSSE correction ensures that all components of an association process are treated in the same basis set [52,53,54]. BSSE treatment was recently applied to various fullerenic species [55,56,57,58].
The DFT approach can be checked using a more advanced second-order Møller–Plesset perturbation treatment (MP2) [59]. There is also another more advanced technique—B2PLYP(D) treatment [60]—that combines the MP2 and DFT methods.
The MPWB1K and MP2 calculations [30] show a substantial encapsulation energy for H2@C60, i.e., the potential energy change during the reaction
H 2 ( g ) + C 60 ( g ) = H 2 @ C 60 ( g ) .
The energy difference between the two methods is 4 kcal/mol (6–31G** basis, without the BSSE term). However, when moving to the still larger basis sets 6–311G(2d,2p) and (d,p)-6–311G**, the two methods differ by only 1 kcal/mol. Moreover, calculations with the SCS MP2 approach [61] reduce the energy gain by about 2 kcal/mol. Overall, it can be concluded [30] that the storage of H2 in the C60 cage is associated with a gain of about −4 kcal/mol in potential energy.
The influence of molecular structure optimizations for H2@C60 was also studied in [30]. It was found that, after full molecular geometry optimization was performed with the MPWB1K/6–31G** approach, the change in the energy reduction was only about 0.44 kcal/mol. The cage C-C bonds were changed by the encapsulation only at the level of the fourth digit. The cage 5/6-type C-C bonds optimized by the MPWB1K/6–31G** approach varied from 1.4377 Å to 1.4388 Å and the cage 6/6-type bonds varied between 1.3792 Å and 1.3798 Å. Let us note that orientation of the H2 unit towards two six-membered rings (and not towards two five-membered rings) brings about a change in the energy of just 0.1 kcal/mol. In other words, H2 can nearly freely rotate in the cage [62] (this feature also appears for some metallofullerenes [63]).

3. N2 Encapsulation in C60

The endohedral system was treated [26,28] with the B3LYP and PW91 functionals [64,65] in the 3–21G basis set. The lowest energy structure of N2@C60 exhibited orientation towards a pair of parallel five-membered rings; the endohedral belonged to the D 5 d point group of symmetry (denoted as a 5:5 structure [28]). The subsequent vibrational treatment confirmed that the N2@C60 5:5 structure had no imaginary vibrational frequencies, i.e., it was indeed a local energy minimum. For N2@C60, the encapsulation energies defined by the process
N 2 ( g ) + C 60 ( g ) = N 2 @ C 60 ( g ) .
had [26], at the MP2=FC/6–31G* level without and with BSSE correction (in the PW91/3–21G optimized geometry), substantial values of −17.5 and −9.28 kcal/mol, respectively, i.e., about two times higher potential energy gain compared to H2@C60.
The hypersurface of N2@C60 interaction energy is rather shallow. Consider a related N2@C60 stationary point that has its N2 molecule pointing towards a pair of parallel six-membered rings, denoted as 6:6 species. However, the 6:6 species is actually a saddle point (i.e., not a local energy minimum). Nevertheless, the 6:6 structure is still very close in energy to the 5:5 orientation—the calculations place the 6:6 structure less than 0.1 kcal/mol above the 5:5 species. This energy difference shows that the surface of interaction energy is flat. Consequently, encapsulated species should exhibit vibrational motion with large amplitudes at elevated temperatures. For the 5:5 N2@C60 optimized geometry, the shortest N-C distance calculated with the B3LYP/3–21G and PW91/3–21G approach is 3.046 and 3.049 Å, respectively. The optimized geometries themselves are not particularly sensitive to the selection of the DFT functional (in contrast to the encapsulation energetics). We also mention that the Mulliken atomic charge calculated on the nitrogen atoms is small—about 0.01 in elementary charge units—meaning that the charge transfer to the cage is negligible.
N2@C60 has 180 vibrational modes: one is the N-N bond stretching mode, five modes are vibrations of the N2 unit against the cage, and the remaining 174 modes are skeletal vibrations of the C60 cage. The N-N bond stretching frequency is only slightly affected by the encapsulation. Let us mention that the N-N stretching frequency increases in the B3LYP/3–21G method but decreases in the PW91/3–21G approach. The shifts in the N-N stretching frequency upon encapsulation are parallel with the calculated changes of the N-N bond length as in the B3LYP/3–21G approach, the bond is slightly shorter, while in the PW91/3–21G method, the N-N bond is longer by 0.0003 Å. The five frequencies for the N2-cage vibrations have rather low values, being similar to those of the La@C82 species [66].
The high icosahedral symmetry of C60 considerably reduces [67,68,69] its infrared (IR) and Raman spectra, e.g., just four T 1 u three-times degenerate vibrational species are actually present in its IR spectrum. However, when the icosahedral symmetry is reduced by encapsulation of N2, the spectral symmetry rules are, strictly speaking, different. Nevertheless, for N2@C60, just twelve cage modes exhibit substantial IR intensities. These vibrational modes active in the IR spectrum actually originate from the four T 1 u (three-times degenerate) modes of the pristine C60. Such degeneracy removal should be seen in the experimental vibrational spectra of N2@C60 once recorded. Similar features can, for example, be seen in the vibrational spectra [46] of CH4@C60 (Figure 1).
The encapsulation energy for Equation (2) is the energy change during the association process. The corresponding enthalpy term for a temperature T, Δ H T o , is obtained by means of the ZPE (zero-point) vibrational energy and related heat-content terms. When the corresponding entropy term Δ S T o is calculated [70], we can arrive at the Gibbs energy change Δ G T o that describes the equilibrium thermodynamics. With the molecular partition functions based on the PW91/3–21G characteristics, the T Δ S T o contribution is equal to −5.95 kcal/mol at room temperature.
Let us note that the entropy evaluations need a careful approach to the rotational symmetry numbers, as some quantum-chemical program packages do not work with the proper symmetry number [69,71] (namely 60) of the icosahedral C60. Moreover, another important issue is the averaging effect of the encapsulate motions in the cage. If the encapsulate moves almost freely inside the C60 cage, it can recover the icosahedral symmetry for the endohedral. Thus, there are two ways [26,44,72,73] to reflect internal motions in the value of the endohedral symmetry number. One approach uses the static (non-icosahedral) symmetry of a rigid endohedral species, while the other considers the icosahedral symmetry produced by the fluctional behaviour [72,73] of the encapsulate.
When the entropy contribution T Δ S T o is used with the enthalpy Δ H T o term based on the MP2 = FC/6–31G* encapsulation energy corrected for the BSSE error, the final Gibbs energy change Δ G T o for reaction (2) at room temperature amounts [26] to −2.64 kcal/mol (the value corresponds to the term, for example, for the water dimerization in the gas phase [74]). The Gibbs energy term represents the driving thermodynamic force for the encapsulation process (2), described by the presently available computational data.

4. NH3 Encapsulation in C60

While N2@C60 is relatively well known based on experiments [2,3], a related endohedral NH3@C60 is yet to be prepared and has been characterized only by calculations [26,75,76], in particular by the energy gain for the encapsulation process:
NH 3 ( g ) + C 60 ( g ) = NH 3 @ C 60 ( g ) .
The encapsulation energy calculated at the MP2 = FC/6–31G** level with the BSSE correction amounts [26] to −5.23 kcal/mol (i.e., about one half of the above energy gain with N2@C60, or comparable to the energy gain [77] in the N@C60 formation). Interestingly, the energy gain is slightly larger in the MP2 = FC/6–31G** approach (−17.89 kcal/mol without BSSE) than from the MP2 = FC/6–31G* method (−16.56 kcal/mol without BSSE). The influence of the DFT functional itself on the molecular structure optimization is again rather insignificant. For NH3@C60, calculations with the B3LYP/3–21G and PW91/3–21G geometries produce encapsulation energies of −5.23 and −5.25 kcal/mol, respectively. Calculations at room temperature for NH3@C60 yield an entropy term T Δ S T o of −5.46 kcal/mol. The Gibbs energy change Δ G T o for reaction (3) is [26] 1.53 kcal/mol, indicating lower stability of NH3@C60 compared to N2@C60.
Let us add for the sake of completeness that the Δ G T o changes (or equilibrium constants of encapsulation) should be considered [78] with the partial pressures of reaction components in experimental conditions. Moreover, catalytic [79,80] and kinetic [14,81,82,83] issues are also potentially involved.

5. CO Encapsulation in C60

CO@C60 was prepared [2,13] by heating under high pressure and placed [20,21] inside open-cage C60 derivatives. Evaluation [40] of the encapsulation process
CO ( g ) + C 60 ( g ) = CO @ C 60 ( g )
for the 5:5 structure optimized at the MPWB1K/6–31G* level with the MP2 = FU/6–311 + G* approach including the BSSE correction gives an encapsulation energy of −12.5 kcal/mol. The thermodynamic treatment can produce for process (4) the encapsulation equilibrium constant K p , e n c , T , exhibiting a relatively fast temperature decrease. Nevertheless, under the synthetic conditions applied [2,13] for CO@C60 preparation (650 °C; 3000 atm), the relative CO@C60 fraction
p C O @ C 60 p C 60 = p C O K p , e n c , T
should be [40] about 3.5%. The evaluation yields an upper limit—the equilibrium should be enabled by a convenient kinetics that in the cage produces a temporary window [2,7,81,82,83,84,85]. In addition, a catalytic action can be required [13,81,85,86,87]. In comparison with the high-pressure and high-temperature encapsulations of water [36,37,38,42], the pressure of carbon monoxide is not controlled by its saturation regime (as the critical temperature of CO is [88] approximately 132 K), meaning that any pressure could in principle be considered. However, there should still be limitations on temperature in order not to allow, e.g., CO dissociation and the production of side derivatives, such as C60O [89,90,91]. An interesting issue concerning the smallest carbon cage [92,93] enabling the encapsulation of CO may also arise. This problem can be handled by means of topological techniques [94,95,96,97].
The IR spectrum of CO@C60 was simulated at the MPWB1K/3–21G level [40]. The stretching mode for encapsulated CO has a frequency of 2150 cm−1, which is 38 cm−1 smaller than at the same level calculated for gas-phase carbon monoxide. For CO encapsulated in derivatives of open C60 [20,21], vibrational frequencies were reported at 2125, 2118, and 2112 cm−1. These frequencies are smaller than the experimental fundamental of gas-phase CO [98] (2143 cm−1, shifting compared to the free CO by 18–31 cm−1). Thus, agreement between the observed and calculated frequency shifts is reasonable, considering differences in the treated terms (anharmonic vs. harmonic frequencies, open vs. closed cages).
Observed shifts in 13C NMR spectra are also reported [20,21] for CO placed in derivatives of open C60. 13C NMR shifts were evaluated with the MPWB1K/6–311 + G* method (in the MPWB1K/6–31G* calculated molecular structure). The experimental NMR shifts [20,21] are smaller by about 10 ppm than those observed for carbon monoxide in solution, while this reduction for the calculated shifts [40] is about 9.7 ppm.
The dipole moment for free CO [99] is rather small (0.122 D), and both experiments [99] and calculations [100,101,102] qualitatively conclude its charge distribution as CO+, i.e., different from polarity derived [103,104] from electronegativity reasoning (however, the calculated term depends on the applied method). Interestingly, when CO@C60 is calculated with the MPWB1K/6–311 + G* approach, the charge on the oxygen atom is 0.92 and the charge on the carbon in CO is 0.28. The MPWB1K/6–311 + G*-calculated value of the endohedral dipole moment is only 0.04 D. The more advanced MP2 = FU/6–311 + G* method leads to a somewhat different picture—the charge on the O atom 1.55, that on the C atom is −1.02, and the system dipole moment is 0.128 D. Another issue is the change in polarizability after encapsulation [105,106], which can also be illustrated with the MPWB1K/3–21G method. For CO@C60, the isotropic polarizability is 62.33 Å3, for C60, it is 62.29 Å3, for CO, it is 1.22 Å3, and the relative reduction owing to encapsulation is equal to −1.18 Å3, in agreement with findings [106] for other fullerene-containing molecules.

6. H2O2 Encapsulation in C60

Recently, H2O2 was placed [22,23] inside a derivative with open-cage C60 at room temperature and atmospheric pressure. The observations were followed by model calculations [43] of the H2O2@C60 endohedral, describing the encapsulation
H 2 O 2 ( g ) + C 60 ( g ) = H 2 O 2 @ C 60 ( g ) .
The M06-2X/6–31++G** molecular structure calculations give two isomers, A and B [43], among which the B isomer has a slightly higher energy, though only by 0.05 kcal/mol. The structural features of the encapsulated H2O2 molecule are not very different from those of free H2O2. The HOOH torsion angle is reduced by about 14° owing to the encapsulation. The shortest distances of the O and H atoms from the cage carbons are still in the range of non-bonding interactions. The calculated rotational constants of the H2O2@C60 isomers have basically identical values, meaning that the species would not be distinguished in rotational spectra.
The harmonic IR spectrum of H2O2@C60(A) was simulated [43] at the M06-2X/6–31++G** level. The free H2O2 possesses six normal modes of vibration; however, only four have substantial IR intensities in the M06-2X/6–31++G** evaluations, viz. 395 cm−1 (internal rotation or torsion mode [107]), 1339 cm−1 (bond-angle deformation), 3849 cm−1 (asymmetric O-H bond stretching), and 3849 cm−1 (symmetric O-H bond stretching). The vibrational modes exhibiting low IR intensity are 1038 cm−1 (O-O bond stretching) and 1478 cm−1 (bond-angle deformation). The four modes with higher IR intensity are basically also seen in H2O2@C60(A) (with some shifts in frequencies): 434 cm−1 (internal rotation or torsion), 1388 cm−1 (bond-angle deformation), 3806 cm−1 (asymmetric O-H bond stretching), and 3821 cm−1 (symmetric O-H bond stretching). The two H2O2 vibrational modes of lower intensity in the IR spectrum have endohedral frequencies of 1078 cm−1 (O-O bond stretching) and 1460.3 cm−1 (bond-angle deformation). Overall, the vibrational spectrum could help in H2O2@C60 laboratory identification once it is synthesized, and could even assist in the search for non-metallic fullerene endohedrals in the interstellar spectra (free hydrogen peroxide is in fact known to be present [108] in the interstellar space).
The charge distribution in hydrogen peroxide is not influenced significantly by encapsulation. In free H2O2, the M06-2X/3–21G-calculated charge on hydrogens is 0.387, while on oxygens, it is equal to −0.387. In H2O2@C60(A), the calculated charge on hydrogens reaches 0.411 on average, and on oxygens, the average is −0.381. Thus, there is quite a small negative charge transfer to the cage. The most negative charge on carbons is −0.0365 (with this particular carbon being located close to the hydrogen). On the other hand, the most positive charge on carbons reaches 0.0133 (this carbon is located close to the oxygen).
The potential energy gains in the formation of H2O2@C60(A) produced via process (6) are calculated [43] as −12.4 and −12.1 kcal/mol with the BSSE-corrected MP2 = FU/6–311++G** and B2PLYPD = FU/6–311++G** treatments, respectively. Thus, both of the considered advanced treatments, MP2 and B2PLYPD, produce almost identical results. The energy term differs by about 1 kcal/mol with the 6–311++G** and 6–31++G** standard basis sets (which agrees with the calculations [30] for H2@C60). The encapsulation energy is comparable to the values found [40] for CO@C60 as well as [26,28,30] for N2@C60. However, the gain in the potential energy is lower than that calculated [42] for (H2O)2@C70 (−18.4 kcal/mol) or for (H2O)2@ D 2 (22)-C84 (−17.4 kcal/mol).
Let us note that hydrogen peroxide has a critical temperature rather similar to that of water [88]. However, there are temperature restrictions to avoid hydrogen peroxide decomposition (and the formation of cage derivatives). Overall, a high-pressure and high-temperature preparation [2,10,11,12,13,14] of H2O2@C60 could in principle be possible.

7. Encapsulation of H2O and Its Aggregates into Fullerenes

Compared to other non-metallic endohedral fullerenes like [2,13] N2@C60 or CO@C60, the water dimer would be rather large to fit in the C60 cage (though H2O@C60 can be prepared [18]). Figure 2 presents [36] the geometrical structure of (H2O)2@C60 optimized with the M06-2X/6–31G** method. The encapsulation energy of the water monomer calculated with the MP2/6–31G approach (without the BSSE term) is [27] −9.9 kcal/mol. However, the inclusion of the water dimer into C60 is repulsive at the same level, i.e., it exhibits [27] a positive (and not negative) encapsulation energy of 24.5 kcal/mol. The calculations instead indicated [36] C84 cages as convenient fullerenes for the water-dimer encapsulation, particular its two most stable species [109,110,111], conventionally labelled as D 2 (22)-C84 and D 2 d (23)-C84 (coded by their symmetries and serial enumeration numbers).
Beyond the previously discussed encapsulations, the C84 cages are represented [109,110,111,112,113,114,115,116,117,118,119] by twenty four [116] C84 isomers that obey the isolated pentagon rule (or IPR). However, the D 2 (22)-C84 and D 2 d (23)-C84 species represent the most populated isomers (though some minor species are reported by Dennis et al. [119]). Previously, based on their NMR observations, Kikuchi et al. [114] concluded that C84 has two major isomers, namely of D 2 and D 2 d symmetries observed in populations of 2:1. According to the semiempirical MNDO calculations [117], the D 2 d cage represents the lowest energy isomer (in other words, the global or lowest energy minimum), though the D 2 d species is placed just 0.5 kcal/mol below the D 2 cage. It should be noted that in the more sophisticated M06-2X/6–31G** approach [37], the D 2 d cage is placed only 0.3 kcal/mol below the D 2 species. In fact, the set of C84 isomers is the first reported case in which the observed most populated fullerene cage does not represent [72,120] the lowest energy isomer (such stability interchanges are also known for metallofullerenes [121,122,123,124,125]). C84 was calculated [109] as an isomeric set consisting of twenty four local energy minima—their energetics, molecular structures, and harmonic vibrational frequencies were produced by the semiempirical MNDO method [117]. The D 2 d isomer is the most populated species in the equilibrium mixture only up to a temperature of 276 K, while beyond this temperature, it is replaced by the D 2 isomer [109]. For example, at a temperature of 1000 K, the D 2 (22)-C84 and D 2 d (23)-C84 cage represents 60.3% and 34.2%, respectively, of the equilibrium 24-member isomeric mixture. In fact, the D 2 structure is chiral, which represents an important contribution [110] to the relative stability interchange. The two most populated C84 isomers can be isolated by chromatography [118] and belong to the most common higher fullerenes.
The calculations in [37] deal with a set of three gas-phase equilibrium processes:
2 H 2 O ( g ) = ( H 2 O ) 2 ( g )
( H 2 O ) 2 ( g ) + D 2 ( 22 ) C 84 ( g ) = ( H 2 O ) 2 @ D 2 ( 22 ) C 84 ( g )
( H 2 O ) 2 ( g ) + D 2 d ( 23 ) C 84 ( g ) = ( H 2 O ) 2 @ D 2 d ( 23 ) C 84 ( g )
The water-dimerization enthalpy at a temperature of absolute zero Δ H 0 , 2 o calculated [37] by the G3&MP2 = (Full)/AUG-cc-pVQZ method amounts to −3.255 kcal/mol. This value is almost the same as the recent spectroscopic term [126,127] of −3.159 ± 0.029 kcal/mol.
The M06-2X/6–31++G** energy gain in the water-dimer encapsulation inside the D 2 (22)-C84 cage is calculated [37,42] as −17.4 kcal/mol, while for the D 2 d (23)-C84 cage, it is −14.4 kcal/mol. The terms include the BSSE correction and even the so-called steric correction [56]. The energy gain from various approaches [37] is consistently larger for the (H2O)2@ D 2 (22)-C84 encapsulation. The calculated data [37] lead to the finding that encapsulation of (H2O)2 into both studied C84 cages should be connected with a significant energy gain. This finding (when considered with the calculated increase [37,128] of the water-dimer population in the saturated water vapor with temperature) indicates the C84 isomers as further potential targets for the experimental investigation of fullerenes with the encapsulated water dimers.
The above-mentioned steric correction [56] basically reflects the cage distortion. First, the complete geometry optimization of the complex species (H2O)2@C84 is performed for both (H2O)2@ D 2 (22)-C84 and (H2O)2@ D 2 d (23)-C84 endohedrals. Traditionally, the BSSE treatment deals with the geometry of the components straightforwardly taken over from the already optimized complex aggregate. However, the steric-corrected BSSE treatment [56] deals with the corresponding optimized empty C84 cage. Such steric correction is also evaluated for the water dimer. Interestingly, the computed O-O distances (2.745 and 2.680 Å in the D 2 and D 2 d cage, respectively) are somewhat shorter compared to the observed value [129,130] of 2.98 ± 0.04 Å in the free (H2O)2.
The encapsulation equilibrium constants for processes (8) and (9) can also be evaluated [39]. The thermodynamic treatment [39] shows that the population ratio of (H2O)2@ D 2 (22)-C84 and (H2O)2@ D 2 d (23)-C84 decreases with increasing temperature, approaching 2:1 for elevated temperatures.
Finally, simultaneous encapsulations of H2O and (H2O)2 in the D 2 (22)-C84 cage are calculated in ref. [38]. In the M06-2X/6–31++G** approach, the monomer encapsulation in D 2 (22)-C84 provides a gain in energy of about −10.7 kcal/mol. The ratio between (H2O)2@C84 and H2O@C84 is also evaluated [38] using the encapsulation equilibrium constants and is close to 1:2.
Similar stability calculations can also be carried out [42] for (H2O)3 encapsulation in the D 2 (22)-C84 fullerene cage:
( H 2 O ) 3 ( g ) + D 2 ( 22 ) C 84 ( g ) = ( H 2 O ) 3 @ D 2 ( 22 ) C 84 ( g ) .
For example, if the energy gain is calculated for the cyclic (H2O)3 encapsulation in D 2 (22)-C84 using the M06-2X/6–31++G** approach including the BSSE error, it is concluded that the water trimer encapsulation in C84 leads to the gain in the potential energy of −10.4 kcal/mol. The encapsulated water trimer can exhibit two different organizations: either the structure present in the free gas-phase water trimer (trans,  C 1 point group of symmetry) or the conformation in which three H atoms (not involved in the H-bond) are on the same side of the O-O-O plane (cis,  C 3 point group of symmetry) (Figure 3). The species differ in energy by only 0.071 kcal/mol. Their gas-phase equilibrium
C 3 ( H 2 O ) 3 @ D 2 ( 22 ) C 84 ( g ) = C 1 ( H 2 O ) 3 @ D 2 ( 22 ) C 84 ( g )
should yield comparable concentrations at higher temperatures. The isomeric concentrations can be calculated [131,132] with the M06-2X/6–31++G** structure and vibrational and energy parameters for the construction of the partition functions (in the commonly used rigid rotor and harmonic oscillator approximation). As can be seen [42] in Figure 4, the concentrations become nearly equimolar quite quickly. At room temperature, the species C 3 ( H 2 O ) 3 @ D 2 ( 22 ) C 84 (with lower potential energy) represents 57% of the equilibrium mixture.

8. Encapsulation of H2O and HF in C70

HF and H2O can be encapsulated in the C70 IPR fullerene cage leading to the observed [41] H2O·HF@C70 species:
H 2 O ( g ) + HF ( g ) + C 70 ( g ) = H 2 O · HF @ C 70 ( g ) .
The encapsulation energy for process (12) calculated using the B2PLYPD/6–31++G** method with the CP3 (case of three reaction components) BSSE correction (including the steric term) is equal [45] to −25.8 kcal/mol (when applying the 6–311++G** atomic basis to −26.0 kcal/mol). The encapsulation energy is evaluated in the M06-2X/6–31++G**-optimized structures. In fact, the M06-2X/6–31++G**-calculated structure [45] of the H2O·HF@C70 species agrees reasonably well with the observed characteristics [41]. In particular, the observed hydrogen bond is 1.39 Å long, while its calculated length is 1.481 Å, and the experimental F-O distance of 2.438 Å is also close to the calculated distance of 2.447 Å. It should be noted that the computations deal with a free (gas-phase) H2O·HF@C70, while the X-ray observation [41] deals with a porphyrin co-crystal. Moreover, the observed [133] dissociation energy for the free H2O·HF dimer itself is also well reproduced [45] with the B2PLYPD/6–311++G** method. The obtained equilibrium constant for encapsulation [45] corresponds to the terms calculated [44] for other studied encapsulations. Hence, it is not excluded that H2O·HF@C70 could also be produced by the direct catalytic high-pressure and high-temperature synthesis [2,13].

9. Conclusions

The surveyed cases document that the computational evaluations can fruitfully cooperate with observations of the fullerene endohedrals containing non-metallic encapsulates and thus support and rationalize the experimental findings. The results therefore encourage further such computational studies of nanocarbon endohedrals with encapsulated non-metal species [44,93,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150] such as H2, N2, CO, HF, NH3, H2O2, H2O, and their aggregates. Future studies should further develop predictions of stability and even populations by applying still more advanced quantum-chemical approaches. Moreover, partition functions for the description of temperature effects should also be further refined so that the description can be based not only on the encapsulation potential energy changes but increasingly also on the encapsulation Gibbs energy terms.

Funding

The related research has received support from 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), the MetaCentrum (LM2010005), and CERIT-SC (CZ.1.05/3.2.00/08.0144) computing facilities. The initial phase of this research line was supported by the Alexander von Humboldt-Stiftung and the Max-Planck-Institut für Chemie (Otto-Hahn-Institut).

Acknowledgments

The authors also wish to thank the following organizations for kindly permitting the reprinting of the author’s own copyrighted materials: American Scientific Publishers; Elsevier Scientific Publishing Company.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dresselhaus, M.S.; Dresselhaus, G.; Eklund, P.C. Science of Fullerenes and Carbon Nanotubes: Their Properties and Applications; Academic Press: San Diego, CA, USA, 1996; p. 316. [Google Scholar]
  2. Peres, T.; Cao, B.P.; Cui, W.D.; Khong, A.; Cross, R.J.; Saunders, M.; Lifshitz, C. Some new diatomic molecule containing endohedral fullerenes. Int. J. Mass Spectr. 2001, 210/211, 241–247. [Google Scholar] [CrossRef]
  3. Suetsuna, T.; Dragoe, N.; Harneit, W.; Weidinger, A.; Shimotani, H.; Ito, S.; Takagi, H.; Kitazawa, K. Separation of N2@C60 and N@C60. Chem. Eur. J. 2002, 8, 5079–5083. [Google Scholar] [CrossRef]
  4. Murphy, T.A.; Pawlik, T.; Weidinger, A.; Höhne, M.; Alcala, R.; Spaeth, J.-M. Observation of atomlike nitrogen in nitrogen-implanted solid C60. Phys. Rev. Lett. 1996, 77, 1075–1078. [Google Scholar] [CrossRef]
  5. Knapp, C.; Dinse, K.-P.; Pietzak, B.; Waiblinger, M.; Weidinger, A. Fourier transform EPR study of N@C60 in solution. Chem. Phys. Lett. 1997, 272, 433–437. [Google Scholar] [CrossRef]
  6. Pietzak, B.; Waiblinger, M.; Murphy, T.A.; Weidinger, A.; Höhne, M.; Dietel, E.; Hirsch, A. Buckminsterfullerene C60: A chemical Faraday cage for atomic nitrogen. Chem. Phys. Lett. 1997, 279, 259–263. [Google Scholar] [CrossRef]
  7. Cao, B.P.; Peres, T.; Cross, R.J.; Saunders, M.; Lifshitz, C. Do nitrogen-atom-containing endohedral fullerenes undergo the shrink-wrap mechanism? J. Phys. Chem. A 2001, 105, 2142–2146. [Google Scholar] [CrossRef]
  8. Kobayashi, K.; Nagase, S.; Dinse, K.-P. A theoretical study of spin density distributions and isotropic hyperfine couplings of N and P atoms in N@C60, P@C60, N@C70, N@C60(CH2)6, and N@C60(SiH2)6. Chem. Phys. Lett. 2003, 377, 93–98. [Google Scholar] [CrossRef]
  9. Wakahara, T.; Matsunaga, Y.; Katayama, A.; Maeda, Y.; Kako, M.; Akasaka, T.; Okamura, M.; Kato, T.; Choe, Y.K.; Kobayashi, K.; et al. A comparison of the photochemical reactivity of N@C60 and C60: Photolysis with disilirane. Chem. Commun. 2003, 39, 2940–2941. [Google Scholar] [CrossRef]
  10. Saunders, M.; Jiménez-Vázquez, H.A.; Cross, R.J.; Poreda, R.J. Stable compounds of helium and neon: He@C60 and Ne@C60. Science 1993, 259, 1428–1430. [Google Scholar] [CrossRef]
  11. Saunders, M.; Jiménez-Vázquez, H.A.; Cross, R.J.; Mroczkowski, S.; Freedberg, D.I.; Anet, F.A.L. Probing the interior of fullerenes by 3He NMR spectroscopy of endohedral 3He@C60 and 3He@C70. Nature 1994, 367, 256–258. [Google Scholar] [CrossRef]
  12. Cross, R.J.; Saunders, M.; Prinzbach, H. Putting helium inside dodecahedrane. Org. Lett. 1999, 1, 1479–1481. [Google Scholar] [CrossRef]
  13. Cross, R.J.; Saunders, M. Catalyzed incorporation of noble gases in fullerenes. In Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials, Volume 11—Fullerenes for the New Millennium; Kadish, K.M., Kamat, p.V., Guldi, D., Eds.; The Electrochemical Society: Pennington, NJ, USA, 2001; pp. 298–300. [Google Scholar]
  14. Rubin, Y.; Jarrosson, T.; Wang, G.-W.; Bartberger, M.D.; Houk, K.N.; Schick, G.; Saunders, M.; Cross, R.J. Insertion of helium and molecular hydrogen through the orifice of an open fullerene. Angew. Chem. Int. Ed. Engl. 2001, 40, 1543–1546. [Google Scholar] [CrossRef]
  15. Carravetta, M.; Murata, Y.; Murata, M.; Heinmaa, I.; Stern, R.; Tontcheva, A.; Samoson, A.; Rubin, Y.; Komatsu, K.; Levitt, M.H. Solid-state NMR spectroscopy of molecular hydrogen trapped inside an open-cage fullerene. J. Am. Chem. Soc. 2004, 126, 4092–4093. [Google Scholar] [CrossRef] [PubMed]
  16. Iwamatsu, S.-I.; Uozaki, T.; Kobayashi, K.; Re, S.; Nagase, S.; Murata, S. A bowl-shaped fullerene encapsulates a water into the cage. J. Am. Chem. Soc. 2004, 126, 2668–2669. [Google Scholar] [CrossRef]
  17. Komatsu, K.; Murata, M.; Murata, Y. Encapsulation of molecular hydrogen in fullerene C60 by organic synthesis. Science 2005, 307, 238–240. [Google Scholar] [CrossRef]
  18. Kurotobi, K.; Murata, Y. A single molecule of water encapsulated in fullerene C60. Science 2011, 333, 613–616. [Google Scholar] [CrossRef]
  19. Zhang, R.; Murata, M.; Aharen, T.; Wakamiya, A.; Shimoaka, T.; Hasegawa, T.; Murata, Y. Synthesis of a distinct water dimer inside fullerene C70. Nat. Chem. 2016, 8, 435–441. [Google Scholar] [CrossRef]
  20. Iwamatsu, S.; Stanisky, C.M.; Cross, R.J.; Saunders, M.; Mizorogi, N.; Nagase, S.; Murata, S. Carbon monoxide inside an open-cage fullerene. Angew. Chem. Int. Ed. 2006, 45, 5337–5340. [Google Scholar] [CrossRef]
  21. Shi, L.J.; Yang, D.Z.; Colombo, F.; Yu, Y.M.; Zhang, W.X.; Gan, L.B. Punching a carbon atom of C60 into its own cavity to form an endohedral complex CO@C59O6 under mild conditions. Chem. Eur. J. 2013, 19, 16545–16549. [Google Scholar] [CrossRef] [PubMed]
  22. Li, Y.; Lou, N.; Xu, D.; Pan, C.; Lu, X.; Gan, L. Oxygen-delivery materials: Synthesis of an open-cage fullerene derivative suitable for encapsulation of H2O2 and O2. Angew. Chem. Int. Ed. 2018, 57, 14144–14148. [Google Scholar] [CrossRef] [PubMed]
  23. Gan, L. Molecular containers derived from 60.fullerene through peroxide chemistry. Acc. Chem. Res. 2019, 52, 1793–1801. [Google Scholar] [CrossRef]
  24. Cioslowski, J. Endohedral chemistry: Electronic structures of molecules trapped inside the C60 cage. J. Am. Chem. Soc. 1991, 113, 4139–4141. [Google Scholar] [CrossRef]
  25. Charkin, O.P.; Klimenko, N.M.; Charkin, D.O.; Mebel, A.M. Theoretical study of host-guest interaction in model endohedral fullerenes with tetrahedral molecules and ions of MH4 hydrides inside the C60H36, C60H24, C84, and C60 cages. Russ. J. Inorg. Chem. 2004, 49, 868–880. [Google Scholar]
  26. Slanina, Z.; Uhlík, F.; Adamowicz, L.; Nagase, S. Computing fullerene encapsulation of non-metallic molecules: N2@C60 and NH3@C60. Mol. Simul. 2005, 31, 801–806. [Google Scholar] [CrossRef]
  27. Ramachandran, C.N.; Sathyamurthy, N. Water clusters in a confined nonpolar environment. Chem. Phys. Let. 2005, 410, 348–351. [Google Scholar] [CrossRef]
  28. Slanina, Z.; Nagase, S. A computational characterization of N2@C60. Mol. Phys. 2006, 104, 3167–3171. [Google Scholar] [CrossRef]
  29. Shameema, O.; Ramachandran, C.N.; Sathyamurthy, N. Blue shift in X-H stretching frequency of molecules due to confinement. J. Phys. Chem. A 2006, 110, 2–4. [Google Scholar] [CrossRef] [PubMed]
  30. Slanina, Z.; Pulay, P.; Nagase, S. H2, Ne, and N2 energies of encapsulation into C60 evaluated with the MPWB1K functional. J. Chem. Theory Comput. 2006, 2, 782–785. [Google Scholar] [CrossRef]
  31. Mazurek, A.P.; Sadlej-Sosnowska, N. Is fullerene C60 large enough to host an aromatic molecule? Int. J. Quantum Chem. 2011, 111, 2398–2405. [Google Scholar] [CrossRef]
  32. 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]
  33. Varadwaj, A.; Varadwaj, P.R. Can a single molecule of water be completely isolated within the subnano-space inside the fullerene C60 cage? A Quantum chemical prospective. Chem. Eur. J. 2012, 18, 15345–15360. [Google Scholar] [CrossRef]
  34. Farimani, A.B.; Wu, Y.B.; Aluru, N.R. Rotational motion of a single water molecule in a buckyball. Phys. Chem. Chem. Phys. 2013, 15, 17993–18000. [Google Scholar] [CrossRef]
  35. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  36. Uhlík, F.; Slanina, Z.; Lee, S.-L.; Wang, B.-C.; Adamowicz, L.; Nagase, S. Water-dimer stability and its fullerene encapsulations. J. Comput. Theor. Nanosci. 2015, 12, 959–964. [Google Scholar] [CrossRef]
  37. Slanina, Z.; Uhlík, F.; Lu, X.; Akasaka, T.; Lemke, K.H.; Seward, T.M.; Nagase, S.; Adamowicz, L. Calculations of the water-dimer encapsulations into C84. Fullerenes Nanotub. Carbon Nanostruct. 2016, 24, 1–7. [Google Scholar] [CrossRef]
  38. Slanina, Z.; Uhlík, F.; Nagase, S.; Lu, X.; Akasaka, T.; Adamowicz, L. Computed relative populations of D2(22)-C84 endohedrals with encapsulated monomeric and dimeric water. ChemPhysChem 2016, 17, 1109–1111. [Google Scholar] [CrossRef]
  39. Slanina, Z.; Uhlík, F.; Nagase, S.; Akasaka, T.; Adamowicz, L.; Lu, X. Computational comparison of the water-dimer encapsulations into D2(22)-C84 and D2d(23)-C84. ECS J. Solid State Sci. Technol. 2017, 6, M3113–M3115. [Google Scholar] [CrossRef]
  40. Slanina, Z.; Uhlík, F.; Nagase, S.; Akasaka, T.; Adamowicz, L.; Lu, X. A computational characterization of CO@C60. Fullerenes Nanotubes Carbon Nanostruct. 2017, 25, 624–629. [Google Scholar] [CrossRef]
  41. Zhang, R.; Murata, M.; Wakamiya, A.; Shimoaka, T.; Hasegawa, T.; Murata, Y. Isolation of the simplest hydrated acid. Sci. Adv. 2017, 3, e1602833. [Google Scholar] [CrossRef]
  42. Slanina, Z.; Uhlík, F.; Nagase, S.; Akasaka, T.; Lu, X.; Adamowicz, L. Cyclic water-trimer encapsulation into D2(22)-C84 fullerene. Chem. Phys. Lett. 2018, 695, 245–248. [Google Scholar] [CrossRef]
  43. Slanina, Z.; Uhlík, F.; Adamowicz, L.; Pan, C.; Lu, X. A computational characterization of H2O2@C60. Fullerenes Nanotubes Carbon Nanostruct. 2022, 30, 258–262. [Google Scholar] [CrossRef]
  44. 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]
  45. Slanina, Z.; Uhlík, F.; Lu, X.; Akasaka, T.; Adamowicz, L. H2O.HF@C70: Encapsulation energetics and thermodynamics. Inorganics 2023, 11, 123. [Google Scholar] [CrossRef]
  46. Slanina, Z.; Uhlík, F.; Akasaka, T.; Lu, X.; Adamowicz, L. A Computational characterization of CH4@C60. Inorganics 2024, 12, 64. [Google Scholar] [CrossRef]
  47. Zhao, Y.; Truhlar, D.G. Hybrid meta density functional theory methods for thermochemistry, thermochemical kinetics, and noncovalent interactions: The MPW1B95 and MPWB1K models and comparative assessments for hydrogen bonding and van der Waals interactions. J. Phys. Chem. A 2004, 108, 6908–6918. [Google Scholar] [CrossRef]
  48. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  50. 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. D.01; Gaussian Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  51. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the difference of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  52. Schwenke, D.W.; Truhlar, D.G. Systematic study of basis set superposition errors in the calculated interaction energy of two HF molecules. J. Chem. Phys. 1985, 82, 2418–2426. [Google Scholar] [CrossRef]
  53. van Duijneveldt, F.B.; van Duijneveldt—Van de Rijdt, J.G.C.M.; van Lenthe, J.H. State-of-the-Art in counterpoise theory. Chem. Rev. 1994, 94, 1873–1885. [Google Scholar] [CrossRef]
  54. Simon, S.; Bertran, J.; Sodupe, M. Effect of counterpoise correction on the geometries and vibrational frequencies of hydrogen bonded systems. J. Phys. Chem. A 2001, 105, 4359–4364. [Google Scholar] [CrossRef]
  55. Slanina, Z.; Lee, S.-L.; Adamowicz, L.; Uhlík, F.; Nagase, S. Computed atructure and energetics of La@C60. Int. J. Quantum Chem. 2005, 104, 272–277. [Google Scholar] [CrossRef]
  56. 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. Quantum Chem. 2011, 111, 2712–2718. [Google Scholar] [CrossRef]
  57. 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]
  58. 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. Rel. Mat. 2021, 118, 108510. [Google Scholar] [CrossRef]
  59. Moller, C.; Plesset, M.S. Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46, 618–622. [Google Scholar] [CrossRef]
  60. Schwabe, T.; Grimme, S. Double-hybrid density functionals with long-range dispersion corrections: Higher accuracy and extended applicability. Phys. Chem. Chem. Phys. 2007, 9, 3397–3406. [Google Scholar] [CrossRef]
  61. Grimme, S. Improved second-order Møller-Plesset perturbation theory by separate scaling of parallel- and antiparallel-spin pair correlation energies. J. Chem. Phys. 2003, 118, 9095–9102. [Google Scholar] [CrossRef]
  62. Cross, R.J. Does H2 rotate freely inside fullerenes? J. Phys. Chem. 2001, 105, 6943–6944. [Google Scholar] [CrossRef]
  63. 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]
  64. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  65. Tsuzuki, S.; Lüthi, H.P. Interaction energies of van der Waals and hydrogen bonded systems calculated using density functional theory: Assessing the PW91 model. J. Chem. Phys. 2001, 114, 3949–3957. [Google Scholar] [CrossRef]
  66. Lebedkin, S.; Renker, B.; Heid, R.; Schober, H.; Rietschel, H. A spectroscopic study of M@C82 metallofullerenes: Raman, far-infrared, and neutron scattering results. Appl. Phys. A 1998, 66, 273–280. [Google Scholar] [CrossRef]
  67. Slanina, Z.; Rudziński, J.M.; Togasi, M.; Ōsawa, E. Quantum-chemically supported vibrational analysis of giant molecules: The C60 and C70 clusters. J. Mol. Struct. (Theochem) 1989, 202, 169–176. [Google Scholar] [CrossRef]
  68. Frum, C.I.; Engleman Jr., R.; Hedderich, H.G.; Bernath, P.F.; Lamb, L.D.; Huffman, D.R. The infrared emission spectrum of gas-phase C60 (buckmisterfullerene). Chem. Phys. Lett. 1991, 176, 504–508. [Google Scholar] [CrossRef]
  69. Slanina, Z. Some aspects of mathematical chemistry of equilibrium and rate processes: Steps towards a completely non-empirical computer design of syntheses. J. Mol. Struct. (Theochem) 1989, 185, 217–228. [Google Scholar] [CrossRef]
  70. 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]
  71. Slanina, Z.; Uhlík, F.; François, J.-P.; Ōsawa, E. Thermodynamic data bases and optical isomerism. Croat. Chem. Acta 2000, 73, 1047–1055. [Google Scholar]
  72. 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]
  73. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Adamowicz, L.; Akasaka, T.; Nagase, S. Calculations of metallofullerene yields. J. Comput. Theor. Nanosci. 2011, 8, 2233–2239. [Google Scholar] [CrossRef]
  74. Slanina, Z. A comparative study of the water-dimer gas-phase thermodynamics in the BJH- and MCYL-type flexible potentials. Chem. Phys. 1991, 150, 321–329. [Google Scholar] [CrossRef]
  75. Ganji, M.D.; Mohseni, M.; Goli, O. Modeling complexes of NH3 molecules confined in C60 fullerene. J. Mol. Struct. (Theochem) 2009, 913, 54–57. [Google Scholar] [CrossRef]
  76. Medrek, M.; Pluciński, F.; Mazurek, A.P. Endohedral complexes of fullerene C60 with small convalent molecules (H2O, NH3, H2, 2H2, 3H2, 4H2, O2, O3) in the context of potential drug transporter system. Acta Pol. Pharm. 2013, 70, 659–665. [Google Scholar]
  77. Park, J.M.; Tarakeshwar, P.; Kim, K.S.; Clark, T. Nature of the interaction of paramagnetic atoms (A=4N,4P,3O,3S) with π systems and C60: A theoretical investigation of A… C6H6 and endohedral fullerenes A@C60. J. Chem. Phys. 2002, 116, 10684–10691. [Google Scholar] [CrossRef]
  78. Slanina, Z. Entropy-controlled reactions: An interesting textbook error. J. Chem. Educ. 1991, 68, 474–475. [Google Scholar] [CrossRef]
  79. Walsh, T.R.; Wales, D.J. Relaxation dynamics of C60. J. Chem. Phys. 1998, 109, 6691–6700. [Google Scholar] [CrossRef]
  80. Kobayashi, K.; Nagase, S. Structures and electronic properties of endohedral metallofullerenes: Theory and experiment. In Endofullerenes—A New Family of Carbon Clusters; Akasaka, T., Nagase, S., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2002; pp. 99–119. [Google Scholar]
  81. Eggen, B.R.; Heggie, M.I.; Jungnickel, G.; Latham, C.D.; Jones, R.; Briddon, P.R. Autocatalysis during fullerene growth. Science 1996, 272, 87–89. [Google Scholar] [CrossRef]
  82. Patchkovskii, S.; Thiel, W. How does helium get into buckminsterfullerene. J. Am. Chem. Soc. 1996, 118, 7164–7172. [Google Scholar] [CrossRef]
  83. Patchkovskii, S.; Thiel, W. Radical impurity mechanisms for helium incorporation into buckminsterfullerene. Helv. Chim. Acta 1997, 80, 495–509. [Google Scholar] [CrossRef]
  84. Shimshi, R.; Khong, A.; Jiménez-Vázquez, H.A.; Cross, R.J.; Saunders, M. Release of noble gas atoms from inside fullerenes. Tetrahedron 1996, 52, 5143–5148. [Google Scholar] [CrossRef]
  85. Eggen, B.R.; Heggie, M.I.; Jungnickel, G.; Latham, C.D.; Jones, R.; Briddon, P.R. Carbon atoms catalyse fullerene growth. Fullerene Sci. Technol. 1997, 5, 727–745. [Google Scholar] [CrossRef]
  86. Slanina, Z.; Zhao, X.; Uhlík, F.; Ozawa, M.; Ōsawa, 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]
  87. Slanina, Z.; Zhao, X.; Ozawa, M.; Adamowicz, L.; Ōsawa, E. Computational evaluations of the elemental-catalytical effects on the kinetics of the Stone-Wales isomerizations. In Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials: Vol. 10—Chemistry and Physics of Fullerenes and Carbon Nanomaterials; Kamat, p.V., Guldi, D.M., Kadish, K.M., Eds.; The Electrochemical Society: Pennington, NJ, USA, 2000; pp. 129–141. [Google Scholar]
  88. Haynes, W.M.; Lide, D.R.; Bruno, T.J. (Eds.) CRC Handbook of Chemistry and Physics, 95th ed.; CRC Press: Boca Raton, FL, USA, 2014; pp. 6–83. [Google Scholar]
  89. Slanina, Z.; Uhlík, F.; Stobinski, L.; Lin, H.-M.; Adamowicz, L. Computing narrow nanotubes and their derivatives. Int. J. Nanosci. 2002, 1, 303–312. [Google Scholar] [CrossRef]
  90. Slanina, Z.; Stobinski, L.; Tomasik, P.; Lin, H.-M.; Adamowicz, L. Quantum-chemical evaluations of thermodynamics and kinetics of oxygen additions to narrow nanotubes. J. Nanosci. Nanotech. 2003, 3, 193–198. [Google Scholar] [CrossRef]
  91. Slanina, Z.; Uhlík, F.; Juha, L.; Tanabe, K.; Adamowicz, L.; Ōsawa, E. Computations on C84O: Thermodynamic, kinetic and photochemical stability. J. Mol. Struct. (Theochem) 2004, 684, 129–133. [Google Scholar] [CrossRef]
  92. Sure, R.; Tonner, R.; Schwerdtfeger, P. Systematic study of rare gas atoms encapsulated in small fullerenes using dispersion corrected density functional theory. J. Comput. Chem. 2015, 36, 88–96. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, E.; Gao, Y. H2 + H2O → H4O: Synthesizing hyperhydrogenated water in small-sized fullerenes? J. Phys. Chem. A 2023, 127, 1190–1195. [Google Scholar] [CrossRef]
  94. Díaz-Tendero, S.; Alcamí, M.; Martín, F. Fullerene C50: Sphericity takes over, not strain. Chem. Phys. Lett. 2005, 407, 153–158. [Google Scholar] [CrossRef]
  95. Vukicevic, D.; Cataldo, F.; Ori, O.; Graovac, A. Topological efficiency of C66 fullerene. Chem. Phys. Lett. 2011, 501, 442–445. [Google Scholar] [CrossRef]
  96. Slanina, Z.; Chao, M.-C.; Lee, S.-L.; and Gutman, I. On applicability of the Wiener index to estimate relative stabilities of the higher-fullerene IPR isomers. J. Serb. Chem. Soc. 1997, 62, 211–217. [Google Scholar]
  97. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Ōsawa, E. Geometrical and thermodynamic approaches to the relative stabilities of fullerene isomers. MATCH Commun. Math. Comput. Chem. 2001, 44, 335–348. [Google Scholar]
  98. Mina-Camilde, N.; Manzanares, I.C.; Caballero, J.F. Molecular constants of carbon monoxide at v = 0, 1, 2, and 3: A vibrational spectroscopy experiment in physical chemistry. J. Chem. Educ. 1996, 73, 804–807. [Google Scholar] [CrossRef]
  99. Muenter, J.S. Electric dipole moment of carbon monoxide. J. Mol. Spectrosc. 1975, 55, 490–491. [Google Scholar] [CrossRef]
  100. Billingsley II, F.P.; Krauss, M. Multiconfiguration self-consistent-field calculation of the dipole moment function of CO(X1Σ+). J. Chem. Phys. 1974, 60, 4130–4144. [Google Scholar] [CrossRef]
  101. Scuseria, G.E.; Miller, M.D.; Jensen, F.; Geertsen, J. The dipole moment of carbon monoxide. J. Chem. Phys. 1991, 94, 6660–6663. [Google Scholar] [CrossRef]
  102. Barnes, L.A.; Liu, B.; Lindh, R. Bond length, dipole moment, and harmonic frequency of CO. J. Chem. Phys. 1993, 98, 3972–3977. [Google Scholar] [CrossRef]
  103. Politzer, P.; Kammeyer, C.W.; Bauer, S.J.; Hedges, W.L. Polar properties of carbon monoxide. J. Phys. Chem. 1981, 85, 4057–4060. [Google Scholar] [CrossRef]
  104. Harrison, J.F. Relationship between the charge distribution and dipole moment functions of CO and the Related Molecules CS, SiO, and SiS. J. Phys. Chem. A 2006, 110, 10848–10857. [Google Scholar] [CrossRef] [PubMed]
  105. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Reduced and quenched polarizabilities of interior atoms in molecules. Chem. Sci. 2013, 4, 2349–2356. [Google Scholar] [CrossRef]
  106. Sabirov, D.S. Polarizability as a landmark property for fullerene chemistry and materials science. RSC Adv. 2014, 4, 44996–45028. [Google Scholar] [CrossRef]
  107. Ghosh, D.C. A quest for the origin of barrier to the internal rotation of hydrogen peroxide (H2O2) and fluorine peroxide (F2O2). Int. J. Mol. Sci. 2006, 7, 289–319. [Google Scholar] [CrossRef]
  108. Bergman, P.; Parise, B.; Liseau, R.; Larsson, B.; Olofsson, H.; Menten, K.M.; Güsten, R. Detection of interstellar hydrogen peroxide. Astronom. Astrophys. 2011, 531, L8. [Google Scholar] [CrossRef]
  109. Slanina, Z.; François, J.-P.; Kolb, M.; Bakowies, D.; Thiel, W. Calculated relative stabilities of C84. Fullerene Sci. Technol. 1993, 1, 221–230. [Google Scholar] [CrossRef]
  110. Slanina, Z.; Lee, S.-L. Chirality effects upon populations of C84 species. J. Mol. Struct. (Theochem) 1995, 333, 153–158. [Google Scholar] [CrossRef]
  111. Ōsawa, E.; Ueno, H.; Yoshida, M.; Slanina, Z.; Zhao, X.; Nishiyama, M.; Saito, H. Combined topological and energy analysis of the annealing process in fullerene formation. Stone-Wales interconversion pathways among IPR isomers of higher fullerenes. J. Chem. Soc. Perkin Trans. 1998, 2, 943–950. [Google Scholar] [CrossRef]
  112. Diederich, F.; Ettl, R.; Rubin, Y.; Whetten, R.L.; Beck, R.; Alvarez, M.; Anz, S.; Sensharma, D.; Wudl, F.; Khemani, K.C.; et al. The higher fullerenes: Isolation and characterization of C76, C84, C90, C94, and C70O, an oxide of D5h-C70. Science 1991, 252, 548–551. [Google Scholar] [CrossRef]
  113. Beck, R.D.; John, P.S.; Alvarez, M.M.; Diederich, F.; Whetten, R.L. Resilience of all-carbon molecules C60, C70, and C84: A surface-scattering time-of-flight investigation. J. Phys. Chem. 1991, 95, 8402–8409. [Google Scholar] [CrossRef]
  114. Kikuchi, K.; Nakahara, N.; Honda, M.; Suzuki, S.; Saito, K.; Shiromaru, H.; Yamauchi, K.; Ikemoto, I.; Kuramochi, T.; Hino, S.; et al. Separation, detection and UV/Visible absorption spectra of fullerenes, C76, C78 and C84. Chem. Lett. 1991, 1607–1610. [Google Scholar] [CrossRef]
  115. Raghavachari, K.; Rohlfing, C.M. Structures and vibrational frequencies of carbon molecules (C60, C70, and C84). J. Phys. Chem. 1991, 95, 5768–5773. [Google Scholar] [CrossRef]
  116. Manolopoulos, D.E.; Fowler, P.W. Molecular graphs, point groups, and fullerenes. J. Chem. Phys. 1992, 96, 7603–7614. [Google Scholar] [CrossRef]
  117. Bakowies, D.; Kolb, M.; Thiel, W.; Richard, S.; Ahlrichs, R.; Kappes, M.M. Quantum-chemical study of C84 fullerene isomers. Chem. Phys. Lett. 1992, 200, 411–417. [Google Scholar] [CrossRef]
  118. Dennis, T.J.S.; Kai, T.; Tomiyama, T.; Shinohara, H. Isolation and characterisation of the two major isomers of [84]fullerene (C84). Chem. Commun. 1998, 619–620. [Google Scholar] [CrossRef]
  119. Dennis, T.J.S.; Kai, T.; Asato, K.; Tomiyama, T.; Shinohara, H.; Yoshida, T.; Kobayashi, Y.; Ishiwatari, H.; Miyake, Y.; Kikuchi, K.; et al. Isolation and characterization by 13C NMR spectroscopy of [84]fullerene minor isomers. J. Phys. Chem. A 1999, 103, 8747–8752. [Google Scholar] [CrossRef]
  120. Slanina, Z. Does the global energy minimum always mean also the thermodynamically most stable structure? J. Mol. Struct. (Theochem) 1990, 65, 143–152. [Google Scholar] [CrossRef]
  121. Wakahara, T.; Nikawa, H.; Kikuchi, T.; Nakahodo, T.; Rahman, G.M.A.; Tsuchiya, T.; Maeda, Y.; Akasaka, T.; Yoza, K.; Horn, E.; et al. La@C72 Having a non-IPR carbon cage. J. Am. Chem. Soc. 2006, 128, 14228–14229. [Google Scholar] [CrossRef] [PubMed]
  122. Mercado, B.Q.; Stuart, M.A.; Mackey, M.A.; Pickens, J.E.; Confait, B.S.; Stevenson, S.; Easterling, M.L.; Valencia, R.; Rodríguez-Fortea, A.; Poblet, J.M.; et al. Sc2(μ2-O) trapped in a fullerene cage: The isolation and structural characterization of Sc2(μ2-O)@Cs(6)-C82 and the relevance of the thermal and entropic effects in fullerene isomer selection. J. Am. Chem. Soc. 2010, 132, 12098–12105. [Google Scholar] [CrossRef] [PubMed]
  123. Mercado, B.Q.; Chen, N.; Rodríguez-Fortea, A.; Mackey, M.A.; Stevenson, S.; Echegoyen, L.; Poblet, J.M.; Olmstead, M.M.; Balch, A.L. The shape of the Sc2(μ2-S) unit trapped in C82: Crystallographic, computational, and electrochemical studies of the isomers, Sc2(μ2-S)@Cs(6)-C82 and Sc2(μ2-S)@3v(8)-C82. J. Am. Chem. Soc. 2011, 133, 6752–6760. [Google Scholar] [CrossRef]
  124. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Adamowicz, L.; Akasaka, T.; Nagase, S. Stability computations for isomers of La@Cn (n = 72, 74, 76). Molecules 2012, 17, 13146–13156. [Google Scholar] [CrossRef]
  125. Mulet-Gas, M.; Rodríguez-Fortea, A.; Echegoyen, L.; Poblet, J.M. Relevance of thermal effects in the formation of endohedral metallofullerenes: The case of Gd3N@Cs(39663)-C82 and other related systems. Inorg. Chem. 2013, 52, 1954–1959. [Google Scholar] [CrossRef]
  126. Rocher-Casterline, B.E.; Ch’ng, L.C.; Mollner, A.K.; Reisler, H. Communication: Determination of the bond dissociation energy (D0) of the water dimer, (H2O)2, by velocity map imaging. J. Chem. Phys. 2011, 134, 211101. [Google Scholar] [CrossRef]
  127. Ch’ng, L.C.; Samanta, A.K.; Wang, Y.; Bowman, J.M.; Reisler, H. Experimental and theoretical investigations of the dissociation energy (D0) and dynamics of the water trimer, (H2O)3. J. Phys. Chem. A 2013, 117, 7207–7216. [Google Scholar] [CrossRef]
  128. Slanina, Z.; Uhlík, F.; Lee, S.-L.; Nagase, S. Computational modelling for the clustering degree in the saturated steam and the water-containing complexes in the atmosphere. J. Quant. Spectr. Radiat. Transf. 2006, 97, 415–423. [Google Scholar] [CrossRef]
  129. Mukhopadhyay, A.; Cole, W.T.S.; Saykally, R.J. The water dimer I: Experimental characterization. Chem. Phys. Lett. 2015, 633, 13–26. [Google Scholar] [CrossRef]
  130. Mukhopadhyay, A.; Xantheas, S.S.; Saykally, R.J. The water dimer II: Theoretical investigations. Chem. Phys. Lett. 2018, 700, 163–175. [Google Scholar] [CrossRef]
  131. Slanina, Z. Contemporary Theory of Chemical Isomerism; D. Reidel Publishing Company: Prague, Czech Republic; Dordrecht, The Netherlands, 1986. [Google Scholar]
  132. 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]
  133. Legon, A.C.; Millen, D.J.; North, H.M. Experimental determination of the dissociation energies D0 0 and De of H2O...HF. Chem. Phys. Let. 1987, 135, 303–306. [Google Scholar] [CrossRef]
  134. Murata, M.; Murata, Y.; Komatsu, K. Surgery of fullerenes. Chem. Commun. 2008, 6083–6094. [Google Scholar] [CrossRef]
  135. Morinaka, Y.; Sato, S.; Wakamiya, A.; Nikawa, H.; Mizorogi, N.; Tanabe, F.; Murata, M.; Komatsu, K.; Furukawa, K.; Kato, T.; et al. X-ray observation of a helium atom and placing a nitrogen atom inside He@C60 and He@C70. Nat. Commun. 2013, 4, 1554. [Google Scholar] [CrossRef]
  136. Dodziuk, H.; Ruud, K.; Korona, T.; Demissie, T.B. Chiral recognition by fullerenes: CHFClBr enantiomers in the C82 cage. Phys. Chem. Chem. Phys. 2016, 18, 26057–26068. [Google Scholar] [CrossRef] [PubMed]
  137. Srivastava, A.K.; Pandey, S.K.; Misra, N. Prediction of superalkali@C60 endofullerenes, their enhanced stability and interesting properties. Chem. Phys. Lett. 2016, 655–656, 71–75. [Google Scholar] [CrossRef]
  138. Hashikawa, Y.; Murata, M.; Wakamiya, A.; Murata, Y. Synthesis and properties of endohedral aza[60]fullerenes: H2O@C59N and H2@C59N as their dimers and monomers. J. Am. Chem. Soc. 2016, 138, 4096–4104. [Google Scholar] [CrossRef]
  139. Basiuk, V.A.; Basiuk, E.V. Noncovalent complexes of Ih-C80 fullerene with phthalocyanines. Fuller. Nanotub. Carbon Nanostruct. 2018, 26, 69–75. [Google Scholar] [CrossRef]
  140. Bloodworth, S.; Sitinova, G.; Alom, S.; Vidal, S.; Bacanu, G.R.; Elliott, S.J.; Light, M.E.; Herniman, J.M.; Langley, G.J.; Levitt, M.H.; et al. First synthesis and characterization of CH4@C60. Angew. Chem. Int. Ed. 2019, 58, 5038–5043. [Google Scholar] [CrossRef]
  141. Hashikawa, Y.; Kizaki, K.; Hirose, T.; Murata, Y. An orifice design: Water insertion into C60. RSC Adv. 2020, 10, 40406–40410. [Google Scholar] [CrossRef]
  142. Carrillo-Bohórquez, O.; Valdés, Á.; Prosmiti, R. Encapsulation of a water molecule inside C60 fullerene: The impact of confinement on quantum features. J. Chem. Theory Comput. 2021, 17, 5839–5848. [Google Scholar] [CrossRef]
  143. Pizzagalli, L. First principles molecular dynamics calculations of the mechanical properties of endofullerenes containing noble gas atoms or small molecules. Phys. Chem. Chem. Phys. 2022, 24, 9449–9458. [Google Scholar] [CrossRef] [PubMed]
  144. 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. Coor. Chem. Rev. 2022, 471, 214762. [Google Scholar] [CrossRef]
  145. 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.; Springe: Singapore, 2022; pp. 219–264. [Google Scholar]
  146. Huang, G.; Hasegawa, S.; Hashikawa, Y.; Ide, Y.; Hirose, T.; Murata, Y. An H2O2 molecule stabilized inside open-cage C60 derivatives by a hydroxy stopper. Chem. Eur. J. 2022, 28, e202103836. [Google Scholar] [CrossRef]
  147. Hashikawa, Y.; Fujikawa, N.; Murata, Y. NH3 encapsulation in π-extended fullerenes. J. Am. Chem. Soc. 2022, 144, 23292–23296. [Google Scholar] [CrossRef]
  148. Jia, A.; Huang, H.; Zuo, Z.-F.; Peng, Y.-J. Electronic structure and interaction in CH4@C60: A first-principle investigation. J. Mol. Model. 2022, 28, 179. [Google Scholar] [CrossRef] [PubMed]
  149. Hashikawa, Y.; Murata, Y. Water in fullerenes. Bull. Chem. Soc. Jpn. 2023, 96, 943–967. [Google Scholar] [CrossRef]
  150. Li, Y.B.; Biswas, R.; Kopcha, W.P.; Dubroca, T.; Abella, L.; Sun, Y.; Crichton, R.A.; Rathnam, C.; Yang, L.T.; Yeh, Y.W.; et al. Structurally defined water-soluble metallofullerene derivatives towards biomedical applications. Angew. Chem. Int. Ed. Engl. 2023, 62, e202211704. [Google Scholar] [CrossRef] [PubMed]
Figure 1. M06-2X/6–31++G** computed IR and Raman (top) spectrum [46] of CH4@C60.
Figure 1. M06-2X/6–31++G** computed IR and Raman (top) spectrum [46] of CH4@C60.
Nanomaterials 15 01287 g001
Figure 2. The M06-2X/6–31G** optimized structure [36] of (H2O)2@C60.
Figure 2. The M06-2X/6–31G** optimized structure [36] of (H2O)2@C60.
Nanomaterials 15 01287 g002
Figure 3. The M06-2X/6–31++G**-optimized structures [42] of (H2O)3@ D 2 (22)-C18: left— t r a n s -organization of H atoms not involved in the hydrogen bonds ( C 1 species); right— c i s -organization ( C 3 species).
Figure 3. The M06-2X/6–31++G**-optimized structures [42] of (H2O)3@ D 2 (22)-C18: left— t r a n s -organization of H atoms not involved in the hydrogen bonds ( C 1 species); right— c i s -organization ( C 3 species).
Nanomaterials 15 01287 g003
Figure 4. Relative concentrations [42] of the (H2O)3@ D 2 (22)-C84 isomers computed at the M06-2X/6–31++G** level.
Figure 4. Relative concentrations [42] of the (H2O)3@ D 2 (22)-C84 isomers computed at the M06-2X/6–31++G** level.
Nanomaterials 15 01287 g004
Table 1. Illustrative examples of the calculated encapsulation energies.
Table 1. Illustrative examples of the calculated encapsulation energies.
Encapsulation ProcessEncapsulation Energy [kcal/mol]Ref.
H 2 ( g ) + C 60 ( g ) = H 2 @ C 60 ( g ) −4.2[30]
N 2 ( g ) + C 60 ( g ) = N 2 @ C 60 ( g ) −9.3[26]
CO ( g ) + C 60 ( g ) = CO @ C 60 ( g ) −12.5[40]
NH 3 ( g ) + C 60 ( g ) = NH 3 @ C 60 ( g ) −5.2[26]
H 2 O 2 ( g ) + C 60 ( g ) = H 2 O 2 @ C 60 ( g ) −12.4[43]
CH 4 ( g ) + C 60 ( g ) = CH 4 @ C 60 ( g ) −13.9[46]
( H 2 O ) 2 ( g ) + C 70 ( g ) = ( H 2 O ) 2 @ C 70 ( g ) −18.4[42]
( H 2 O ) 2 ( g ) + C 84 ( g ) = ( H 2 O ) 2 @ C 84 ( g ) −17.4[42] a
( H 2 O ) 3 ( g ) + C 84 ( g ) = ( H 2 O ) 3 @ C 84 ( g ) −10.4[42] a
H 2 O ( g ) + HF ( g ) + C 70 ( g ) = H 2 O · HF @ C 70 ( g ) −26.0[45]
a  D 2 ( 22 ) - C 84 .
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. Theoretical Studies of Non-Metal Endohedral Fullerenes. Nanomaterials 2025, 15, 1287. https://doi.org/10.3390/nano15161287

AMA Style

Slanina Z, Uhlík F, Akasaka T, Lu X, Adamowicz L. Theoretical Studies of Non-Metal Endohedral Fullerenes. Nanomaterials. 2025; 15(16):1287. https://doi.org/10.3390/nano15161287

Chicago/Turabian Style

Slanina, Zdeněk, Filip Uhlík, Takeshi Akasaka, Xing Lu, and Ludwik Adamowicz. 2025. "Theoretical Studies of Non-Metal Endohedral Fullerenes" Nanomaterials 15, no. 16: 1287. https://doi.org/10.3390/nano15161287

APA Style

Slanina, Z., Uhlík, F., Akasaka, T., Lu, X., & Adamowicz, L. (2025). Theoretical Studies of Non-Metal Endohedral Fullerenes. Nanomaterials, 15(16), 1287. https://doi.org/10.3390/nano15161287

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop