Next Article in Journal
Cytotoxic Peptidic Metabolites Isolated from the Soil-Derived Fungus Trichoderma atroviride
Previous Article in Journal
The Chelating Abilities of Tertiary Amines with N-O-Donors Towards Cu(II) Ions and the Catalytic Properties of the Resulting Complexes
Previous Article in Special Issue
A Rapid Intelligent Screening of a Three-Band Index for Estimating Soil Copper Content
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Calculations of Structure and IR Spectra of M@C60 and M2@C60 Endofullerenes (M=Sc and Y)

by
Alexey V. Krisilov
1,
Igor V. Nechaev
2,
Vladislav E. Chernov
1,* and
Gie Eli Kallu
2
1
Faculty of Physics, Voronezh State University, 394018 Voronezh, Russia
2
Faculty of Chemistry, Voronezh State University, 394018 Voronezh, Russia
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(16), 3421; https://doi.org/10.3390/molecules30163421
Submission received: 14 July 2025 / Revised: 14 August 2025 / Accepted: 17 August 2025 / Published: 19 August 2025
(This article belongs to the Special Issue Vibrational Spectroscopy and Imaging for Chemical Application)

Abstract

The endohedral metallofullerenes with a rare-earth metal encapsulated into the carbon cage are nanoparticles with potentially wide applications. We present the results of our quantum-chemical modelling of Sc@ C 60 , Y@ C 60 and Sc2@ C 60 , Y2@ C 60 endofullerenes and calculate their structures and vibrational spectra. Our calculations show that the encapsulation of an additional metal atom inside the carbon cage significantly changes the vibrational spectrum of endofullerene. The most significant changes in the far-IR (below 600 cm−1) spectra are due to the metal–carbon cage vibration modes.

1. Introduction

The unique properties of fullerenes offer a wide range of opportunities for fabricating various single-molecule electronic devices [1]. An endohedral fullerene (or endofullerene) is a carbon shell of fullerene C60 or higher fullerenes such as C70, C78 or C82, within which one or more atoms are encapsulated [2,3]. Endohedral metallofullerenes (EMFs) have attracted considerable attention due to their unique properties and wide range of possible applications [4,5,6,7,8,9,10,11]. In addition to their unique electronic properties, endofullerenes have enhanced chemical stability compared to their non-metal-enriched counterparts [12]. EMFs are characterized by large light absorption intensity in the visible and IR ranges, which leads to increased photocurrent and photovoltaic device efficiency [13]. EMFs are charge transfer complexes because they combine both electron donor and electron acceptor in one molecule [14,15]. This property is very important for the operation of photovoltaic systems, where rapid separation of electron–hole pairs is required to prevent recombination and ensure efficient energy conversion. By choosing the type and number of metal atoms embedded in the fullerene, the width of the forbidden band gap can be adjusted, which can optimize the absorption spectrum and increase efficiency of energy conversion in photovoltaic systems [16,17]. Controlling the properties of EMFs by selecting suitable metal atoms and functionalizing the fullerene shell offers great potential for the development of new materials, electronic and optoelectronic devices, including semiconductors, photodetectors, and transistors [18,19,20,21,22].
The geometrical and electronic structure as well as spin properties of scandium and yttrium dimetallofullerenes (di-EMFs) are being actively studied [23,24]. Some EMFs containing scandium and yttrium atoms are chemically stable molecular nanomagnets, which are also known as single-molecule magnets [25]. Structures possessing magnetic bistability can be used to store magnetic information at the molecular level, which attracts researchers in the field of quantum information technology and the development of high-density information storage. Studies of the electronic structure, magnetic interactions and molecular vibrations of EMFs, as well as their dependence on the spin state are of particular interest [26,27]. Scandium and yttrium EMFs also exhibit high efficiency in photovoltaic applications and offer new opportunities to enhance the performance of various optoelectronic devices [28,29,30].
DFT calculations are widely used for modelling the endohedral fullerenes, both with a single atom [31] or a molecule [32] encapsulated inside the carbon cage. In our previous works (see, e.g., [33,34] and references therein) we conducted DFT analysis to compare the structures, symmetries, spins and dipole moments of lanthanide EMFs by means of quantum-chemical calculations and obtained their vibrational spectra in different spin states. In this work, we calculate the stable structures and IR spectra of the mono- and di-EMFs of scandium and yttrium using the density functional method (DFT).
It should be noted that the EMFs of rare-earth metals have already been extracted and studied experimentally. For instance, Ref. [35] studied the La@C60 and Gd@C60 EMFs and the Gd @ C 60 ( CF 3 ) 3 , Gd @ C 60 ( CF 3 ) 5 and La @ C 60 ( CF 3 ) 5 functionalized EMFs. Refs. [36,37,38,39,40] report on synthesis of EMFs with C60 that encapsulate Y, La, Ce, Pr, Nd, Gd atoms. The extraction of these EMFs was performed using aromatic solvents such as pyridine and aniline, the isolation of pure M@C60 EMFs being significantly complicated due to the formation of exohedral adducts with solvent molecules.
During the synthesis of EMFs, structures can be formed with both metal clusters Mx@C2n+2 and metal carbides inside the carbon cage (MxC2)@C2n. Determining whether a carbide or a metal cluster is inside a fullerene has been an important issue since 2001, when metal-carbide EMFs (Sc2C2)@C84 were first isolated [41], up to recent studies of the Y 2 @ C 84 and Y 2 C 2 @ C 82 [42], Sc 2 @ C 70 and Sc 2 @ C 68 [43], Sc 2 C 2 @ C 80 and Sc 2 @ C 82 [44], or Er 2 @ C 80 and Er 2 C 2 @ C 78 [45].
Di-EMFs M2@C80 (M = Sc, Y, Gd, Er, Lu) have a triplet electronic state due to the half-filled low-lying binding orbital of the metal dimer (one unpaired electron is transferred from this orbital to the carbon cage). Such high spin states makes di-EMFs promising candidates for molecular qubits or spin-filters. IR-active metal–cage and metal–metal vibrational modes couple to these spin states via spin–phonon interactions, directly influencing decoherence times. This makes di-EMFs the simplest organometallic compounds with M–M bonds which possess a unique molecular magnetism [46]. Encapsulating the second rare-earth atom into the carbonic cage allows to form single-electron metal–metal bonds which can enhance magnetic coupling between the atoms and create a better-performing single-molecular magnet [27].
In addition, di-EMFs are unique objects in which metal atoms are isolated from the external environment by a carbon cage, which is of interest from the point of view of analyzing the nature of the bond between two metal atoms and studying the influence of additional non-metal atoms inside the fullerene on the stability of the molecular structure [47,48].
Although Y@C60 has already been obtained during synthesis [49], to date, the Y2@C60 di-EMF has not been isolated so far in the synthesis of EMFs. The same is the case for mono- and di-EMFs of scandium with C60 cage. It should be noted that structures and isomers not identified in the first synthesis of endofullerenes of the corresponding metal can sometimes be obtained decades later [50]. Therefore, the calculation of IR spectra of the above EMFs can contribute to their detection and identification.
Raman spectroscopy should be mentioned as a valuable technique to use for the identification of molecules using intuitive parameters obtained from experiments, such as line width, peak intensity, and peak position changes. However, IR spectra are also informative for detection and identification of molecular species. Note for instance that it was IR spectrum calculation that allowed for distinguishing between the Sc 2 @ C 84 [44] and hypothetical Sc 2 C 2 @ C 82 di-EMFs [3].
High-accuracy IR predictions can contribute to a library of vibrational fingerprint searches for EMFs, delivering the essential “spectral coordinates” for detecting and characterizing these species even before their macroscopic isolation has been conducted in the laboratory. This is especially important in the cases when Raman measurements are not available. For example, the growing data recorded by the rapidly developing IR astronomy facilities require knowledge of IR spectra for a large number of molecular species, including organic compounds [51]. EMFs are potential emitters contributing to the observed infrared spectra in fullerene-rich planetary nebulae and circumstellar shells [52] and they can be candidates for explaining several unidentified IR lines. The detection of fullerene derivatives (such as mono- and di-EMFs) in planetary nebulae can enrich our knowledge of the chemical evolution of the Universe.
The experimental frequencies of atom vibrations inside the EMF cage are difficult to measure. The influence of the Rayleigh wing of the scattering of the initial radiation line makes it difficult to detect small frequency shifts. Analysis of far-IR spectra can be complicated by the low energy of photons. Therefore the theoretical calculation of the vibration frequencies of metal atoms in EMFs is an important task.
This paper is organized as follows. The calculation scheme is described in Section 3. The results of the calculations are presented in Section 2. The geometry of the calculated structures is described in Section 2.1, and their stability is shortly discussed in Section 2.2. The general discussion of the vibrational spectra is given in Section 2.3. The IR spectra calculated for both the mono- and di-EMFs consist of many lines, most of which are absent in the IR spectra of empty fullerene C60 due to symmetry forbidding. For metal–cage (M–cage) modes, in Section 2.4 we indicate the atomic displacements for both IR-active vibrations and vibrations that do not appear in the IR spectra. Some of these modes are illustrated by short movies in the Supplementary Materials. In particular, for some M–cage modes we explain the difference in IR activity between mono- and di-EMFs. The calculated spectra are discussed in detail in Section 2.5 for the mono-EMFs and in Section 2.6 for the di-EMFs. The comparison of the spectra calculated for different mono- and di-EMFs is given in Section 2.7. The full list of the calculated IR lines is given in the Appendix A; a visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).

2. Results and Discussion

2.1. Structure of M@C60 and M2@C60 Endofullerenes

According to the obtained results of quantum-chemical calculations, metal atoms are attached to the six-membered carbon ring, which is consistent with the data for the Y2@C80 endofullerene [53].
The calculated structures of endofullerenes are shown in Figure 1. The interatomic distances in the calculated structures are given in Table 1, where C1, C2, C3 are the atoms of the six-membered carbon ring to which the metal atom is attached (see Figure 1a). The distance between the metal atoms inside the Sc2@C60 ( C 3 v ) fullerene is larger than for Y2@C60 ( C 3 v ), because as the size of the metal atom increases, the metal–carbon bond length increases and the metal atoms shift closer to the center of the fullerene. For di-EMFs, the metal–carbon bond lengths are the same for both metal atoms. The growth of the metal–carbon bond length limits the maximum possible distance between metal atoms in the case of their attachment to opposite six-membered carbon rings. In the case of C s symmetry in Sc2@C60 the second metal atom is attached not to the opposite six-membered ring, but to the adjacent one, so the distance between the metal atoms is significantly smaller than in Sc2@C60 of C 3 v symmetry. Note that spin unrestricted DFT calculations show that the metal atom in Gd@C60 EMF situated opposite to the center of a six-membered ring bond [35].
The calculated metal–carbon bond lengths are in agreement with X-ray structural data for Sc2@C82 (bond lengths r(Sc–C) = 2.145–2.225 Å [54]) and Y2@C79N (bond lengths r(Y–C) = 2.336–2.523 Å [55]).

2.2. Energy Stability and Availability of M@C60 and M2@C60 Structures

The geometry optimization was performed with different initial arrangements of metal atoms. To calculate the spectroscopic parameters for each EMF, the structures with the lowest energy were selected. The following stable isomers of the considered EMFs have been found during the geometry optimization in our DFT calculations.
For the Sc2@C60 molecule, the lowest electronic state energy (i.e., zero relative energy) is achieved for the C 3 v symmetry isomer. The closest Sc2@C60 ground state (relative energy 0.11 eV) value is found for the isomer with the C s symmetry. For Y2@C60, the lowest ground electronic state (relative energy 0 eV) is achieved for the C 3 v isomer. Although the Y2@C60 isomer with the C s symmetry is close in energy (relative energy 0.20 eV), this isomer is unstable (i.e., has imaginary frequency i66.41 cm−1) and is not considered further. The imaginary frequency indicates that the structure corresponds not to a minimum, but to a saddle point on the potential energy surface. As for mono-EMFs (Sc@C60 and Y@C60), both of them are characterized by the C s symmetry of the lowest stable state. All structures described here are stable, as evidenced by the absence of imaginary frequencies of molecular vibrations.
The C60 fullerene features a relatively small inner cavity (∼0.7 nm in diameter), which imposes severe steric restrictions for encapsulation of atoms. While C60 can host up to three small Li atoms [56] or even LiF molecule [57], it is much harder to encapsulate metal atoms such as Sc (atomic radius ∼1.62 Å) and Y (atomic radius ∼1.80 Å). Larger metals like Y cause steric and strain effects that destabilize the cage–metal complex. At the same time, larger C2n cages (e.g., with 2n = 80–84) can encapsulate two [58,59] or even three Sc atoms [60]. As for the C60 cage, DFT calculations [56] predict that, for instance, the production yield of Lix@C60 EMFs in the high-temperature synthesis should decrease by several orders of magnitude with x increasing from 1 to 3.
In addition to the structure stability, another explanation of difficulties with experimental synthesis of the considered EMFs can be high-temperature synthesis bias towards larger cages. The primary synthesis method for EMFs is arc-discharge or laser vaporization of metal/graphite composites. Under such high-temperature conditions, carbon cages that encapsulate metals typically favor larger cage sizes to ensure enough space in the inner cavity to place the metal atoms and achieve a local minimum of the potential energy surface of the EMF. Formation of pure C60 cages encapsulating Sc or Y atoms can thereby disfavored kinetically and thermodynamically.
It is possible to determine the molecular formula and structure only for those EMFs that can be extracted from arc discharge soot. Some EMFs can be synthesized but not detected if they are not soluble in commonly used solvents. The reason for poor solubility is low kinetic stability, which implies that the “insoluble” EMFs react with other carbon-containing substances formed in the soot, or polymerize. Then these EMFs can be extracted by modifying the synthesis protocol. Ensuring kinetic stability and solubility can be accomplished by electron transfer and/or chemical functionalization [50].

2.3. Vibrational Spectra of the M@C60 and M2@C60 Endofullerenes

The empty C60 fullerene has high I h symmetry, and theoretical calculations [61] show that of the 46 non-degenerate vibrational modes with different frequency, only four modes ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]) are IR-active and only ten modes are Raman, while the remaining “silent” modes appear in neutron scattering experiments.
The vibrational spectra of endohedral fullerene differ significantly from those of empty fullerene.
First, it occurs due to the involvement of metal atoms in the vibrations of the carbon cage. In particular, our calculation shows that the maximal contribution (10–30%) of the metal atoms takes place for the vibrational ν 1 ν 4 modes of M2@C60 di-EMFs and ν 1 ν 3 modes of M@C60 mono-EMFs. The frequency 102 cm−1 of shear vibrations of the Y2 dimer along the axis perpendicular to the Y–Y bond was calculated in Ref. [63] for Y2@C84 endofullerene. Raman measurements performed in the same study (Ref. [63]) for molecules in crystalline form gave crystal field splitting results of 77–95 cm−1 ( T = 100 K) for metal–cage vibrations. These results are in good agreement with our ν 3 , 4 101–103 cm−1 values. The similar frequency calculated in Ref. [63] for Sc2@C84 is 104 cm−1, while the experimental Raman values lie between 84 and 110 cm−1 ( T = 100 K). Our calculated values ν 1 , 4 79–94 cm−1 for Sc2@C60 ( C s ) and ν 1 , 2 82 cm−1 for Sc2@C60 ( C 3 v ) demonstrate overall agreement with the results of Ref. [63].
Second, the symmetry reduction due to the encapsulation of the metal atom leads to the appearance of lines in the IR spectra of endohedral fullerenes in the of 20–1620 cm−1 range. Most of these lines are absent both in the IR and Raman spectra of individual empty C60 fullerene [61,62] molecules, but some of the vibrational modes can occur in the IR and Raman spectra of C60 crystals (e.g., because of symmetry reduction due to crystal field effects and the presence of different isotopologues [64]).
The vibrations of the carbon cage (in both endohedral and empty fullerenes) are divided into radial (most lines with ν < 800 cm−1) and tangential ( ν > 1000 cm−1) modes. While tangential modes preserve the shape of the carbon cage, radial modes change its shape. The IR-active C60 modes with frequencies 527 and 576 cm−1 are associated with radial motion of carbon atoms, while the 1182 and 1429 cm−1 modes are associated with tangential motion of carbon atoms [65,66].
The so-called “pumpkin” modes are accompanied by a deformation of the carbon cage from a oblate spheroid to a prolate one (much like squeezing and releasing a pumpkin or rubber ball). This type of vibration changes the shape of the carbon cage, while preserving (to first order) its overall volume.
For empty C60, all the five “pumpkin” modes are IR-inactive; they appear in Raman spectra and have the same frequency of 272 cm−1 [67]. Unlike the case of empty fullerene, in endohedral fullerenes the degeneracy of these modes is removed. This type of vibration changes the size of the polarized endofullerene cage, which changes the dipole moment and makes the pumpkin vibrational modes of the mono-EMF carbon cage IR-active. The difference between the frequencies of the EMF pumpkin modes and the corresponding C60 frequencies is due to the formation of metal–carbon bonds.
The “silent” mode with a frequency of 403 cm−1 does not appear in either IR or Raman spectra, but it is registered in neutron scattering experiments [67,68]. This vibrational mode is the only tangential mode with a frequency significantly lower than 1000 cm−1. Unlike all other tangential modes, large fragments of the carbon cage are involved in the coordinated motion in this mode, which reduces the vibration frequency. This mode represents counter torsional vibrations of two carbon cage’s hemispheres. Due to the decrease in symmetry upon encapsulation of the metal dimer, this mode occurs IR-active for Sc2@C60 ( C 3 v ), having a frequency of ν 18 = 387.15 cm−1 and a significant intensity (see Figure 2). For the Sc2@C60 ( C s ) isomer, this mode is also IR-active, with a frequency of ν 20 = 395.12 cm−1, but low intensity (see Figure 2). This mode is also IR-active for Y2@C60 ( C 3 v ), with a frequency of ν 16 = 371.3 cm−1 and significant intensity (see Figure 3). For Sc@C60 and Y@C60, this type of vibrations gives several lines of medium and low intensity in the IR spectrum in the range of 394–406 cm−1. For Sc@C60 this mode has a frequency of ν 16 = 398.32 cm−1, but a weak intensity (see Figure 2). For Y@C60 this mode appears in the IR spectrum at a frequency of ν 16 = 394.37 cm−1 (see Figure 3). Activation of the “silent” mode of C60 is observed due to the influence of the encapsulated metal atom or metal dimer.
The “Breathing” mode represents vibrations with all-round compression–extension of the fullerene’s carbon cage. This mode for empty C60 has a frequency of 496 cm−1, this mode is not IR-active and manifests itself in the Raman spectra [62]. Sc@C60 and Y@C60 vibrations with close to isotropic compression–extension of the carbon framework have a frequency of 484–485 cm−1; due to the reduction of symmetry, these vibrations are IR-active, but the corresponding spectral lines are weak. For Sc2@C60 ( C 3 v ) and Y2@C60 ( C 3 v ), these vibrations are symmetrical with respect to the fullerene center and therefore are not IR-active. However, for Sc2@C60 ( C s ), due to the reduced symmetry the “breathing” mode appears in the IR spectrum with a frequency ν 28 = 455.84 cm−1 and a significant intensity (see Figure 2). The encapsulation of scandium dimer leads to the appearance of IR activity in this Raman mode of C60 fullerene.
Of the tangential vibrations, the pentagonal pinch mode should be mentioned. It is caused by the absolutely symmetric stretching of the pentagonal rings of the carbon cage. The corresponding vibrations are not IR-active for all structures considered here, and their frequencies lie in the range 1502–1512 cm−1 (see Table A1 and Table A2). These values are in order of magnitude consistent with the frequency of the 1470 cm−1 pentagonal-pinch Raman active mode of the empty C60 fullerene, which corresponds to the displacement of carbon atoms to the 12 centers of the pentagonal rings along the fullerene surface [67].
The directions of atomic displacement most typical for the structures calculated in this work are shown by arrows in Figure 4 and Figure 5. Additionally, a visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).

2.4. M–Cage Modes in Mono- and di-EMFs

Compared to empty C60 fullerene, the IR spectrum of both mono- and di-EMFs contains significantly more lines. The spectra of di-EMFs with C 3 v symmetry in the ranges 20–-450, 500–800, and 1200–1600 cm−1 have fewer IR-active vibrational modes compared to mono-EMFs. Indeed, according to our calculations of mono-EMFs, the metal atom is not located in the center of the M@C60 molecule, so it cannot have central symmetry. Geometry optimization lead to different EMF isomers with C 3 v and C s symmetry. In M2@C60 with C 3 v symmetry, both the metal atoms are located on the C 3 axis at the same distance from the center of C60 and are bonded to opposite six-membered rings.
Therefore, the di-EMF M2@C60 molecules of C 3 v have a center of symmetry. This higher (as compared to M2@C60) symmetry prohibits some IR lines. For Sc2@C60 of C s symmetry, this effect is less pronounced due to the metal atoms bonding with the six-membered rings that are placed near the C 3 axis (but not exactly on it).
The symmetry considerations explain why a number of M–cage modes in di-EMFs are not IR-active. Our calculations show that the charge density is positive at the incapsulated metal atom locations. In mono-EMF M@C60, therefore, a motion of the metal atom (which always takes place in M–cage modes) results in a change of the positive charge center. In turn, this displacement of positive charge leads to change of the molecule dipole moment that makes the corresponding vibration mode IR-active. On the contrary, in di-EMFs M2@ C 60 , the metal ions move in opposite directions (with respect to the center of symmetry) in all vibrational modes except for shear vibrations. Therefore most of M–cage modes vibrations are IR-inactive for di-EMFs.
Thus, in the IR spectrum of Y2@C60 compared to Y@C60, the intensity of all M-cage modes (in the 20–450 cm−1 range) decreases sharply, with the exception of ν 6 and ν 26 , associated with shear vibrations of the metal dimer along the Y–Y bond. This effect is caused by the symmetrical motion of metal atoms relative to the center of the fullerene, while the center of positive charge does not shift and the vibrations do not change the dipole moment of the molecule and therefore are not IR-active.
Similarly, in the IR spectrum of Sc2@C60 ( C 3 v ) compared to Sc@C60 (in the 20–450 cm−1 range), with the exception of modes associated with shear vibrations of the metal dimer along the Sc–Sc bond. For Sc2@C60 ( C s ), this effect also occurs, but is less pronounced.

2.5. Spectra of Mono-EMFs of Sc and Y

The calculated vibrational IR spectra of mono-EMFs are presented in Figure 6 and Figure 7. The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).

2.5.1. Sc@C60 ( C s )

In the Sc@C60 spectrum, there are seven vibrational modes in the frequency range up to 300 cm−1: one IR-active mode (with a frequency of 55 cm−1), three weak (low-intensity) IR lines, and three IR-inactive modes. Between 300 and 400 cm−1, there are 11 modes, of which 8 are low-intensity IR and 3 are IR-inactive. The range 400–600 cm−1contains 6 IR-active lines (with frequencies 426, 519, 530, 575, 583, 592 cm−1), 16 weak IR lines, and 9 IR-inactive modes (31 modes in total). There are 40 vibrational modes in the 600–800 cm−1 range, of which 4 are IR-active (606, 677, 688, 779 cm−1), 17 weak IR lines, and 19 IR-inactive modes (see Figure 2).
In the 800–1000 cm−1 spectral range, 11 vibrational Sc@C60 modes were found, of which 4 are low-intensity IR and 7 are IR-inactive. In the range 1000–1200 cm−1, there are 13 vibrational modes, of which 4 are IR-active (with frequencies 1001, 1069, 1104, 1181 cm−1), 7 are IR-weak, and 2 are IR-inactive. In addition, 35 vibrational modes fall within the frequency range 1200–1400 cm−1; among them, 10 are IR-active (1204, 1212, 1225, 1237, 1240, 1293, 1298, 1389, 1390, 1399 cm−1), 18 low-intensity IR lines, and 7 IR-inactive modes. In the range 1400–1600 cm−1 19 IR-active lines were found (with frequencies 1403, 1412, 1434, 1449, 1456, 1459, 1467, 1474, 1483, 1485, 1501, 1519, 1542, 1545, 1598, 1601, 1605, 1611, and 1612 cm−1), 8 low-intensity IR lines, and 2 IR-inactive modes (29 modes in total; see Figure 6).

2.5.2. Y@C60 ( C s )

In the Y@C60 spectrum, there are seven vibrational modes in the frequency range up to 300 cm−1: one IR-active (with a frequency of 53 cm−1), three low-intensity IR, and three IR-inactive. Between 300 and 400 cm−1, 12 modes were found, 8 of which are low-intensity IR and 4 are IR-inactive. The range 400–600 cm−1 contains 6 IR-active lines (with frequencies 422, 516, 524, 578, 591, 598 cm−1), 15 low-intensity IR lines, and 10 IR-inactive modes (31 modes in total). There are 40 vibrational modes in the 600–800 cm−1 range; of which 4 are IR-active (618, 668, 686, 775 cm−1), 17 are low-intensity IR, and 19 are IR-inactive (see Figure 7).
In the range 800–1000 cm−1, there are 10 Y@C60 modes: 4 low-intensity IR modes and 6 IR-inactive modes. Between 1000 and 1200 cm−1, there are 14 vibrational modes, of which 6 are IR-active (with frequencies 1001, 1050, 1095, 1098, 1175, 1195 cm−1) and 8 are low-intensity IR lines. Additionally, 36 vibrational modes fall within the frequency range 1200–1400 cm−1; of these, 14 are IR-active (1204, 1223, 1229, 1235, 1237, 1266, 1288, 1297, 1372, 1384, 1385, 1392, 1396, 1400 cm−1), 19 are low-intensity IR, and 3 are non-IR-active. In the range 1400–1600 cm−1 we found 18 IR-active lines (with frequencies 1420, 1436, 1447, 1455, 1458, 1468, 1472, 1480, 1482, 1496, 1521, 1534, 1537, 1591, 1596, 1603, 1606, and 1608 cm−1), 8 low-intensity IR lines, and 1 IR-inactive mode (27 modes in total; see Figure 3).

2.6. Spectra of di-EMFs of Sc and Y

The calculated vibrational IR spectra of di-EMFs are presented in Figure 8, Figure 9 and Figure 10. The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).

2.6.1. Sc2@C60 ( C 3 v )

The vibrational spectrum of Sc2@C60 with C 3 v symmetry is presented in Figure 8. The vibrational modes ν 1 = 81.95 cm−1 and ν 2 = 82.68 cm−1 represent shear vibrations of the Sc2 dimer along the axis perpendicular to the Sc–Sc bond, the fraction of metal atom involvement being 17%. The vibrational modes ν 3 = 89.02 cm−1 and ν 4 = 89.65 cm−1 correspond to torsional vibrations of the Sc2 dimer with respect to the axis passing perpendicularly through the middle of the metal–metal bond (lateral M–cage mode); the fraction of metal atoms involved is 22%. This type of vibrations is not IR-active, since it does not lead to the displacement of the center of positive charge.
Of the pumpkin modes, the largest fraction (4%) of the involvement of Sc atoms is characteristic of the mode ν 5 = 246.35 cm−1 (compression–stretching of the Sc2 dimer along the Sc–Sc bond). In pumpkin modes ν 6 ν 7 with frequencies 257–258 cm−1 the compression–stretching of the carbon cage goes perpendicular to the Sc–Sc bond. Metal atoms are almost not involved; these vibrations do not appear in IR spectra, as well as in the empty C60 fullerene spectrum. In pumpkin modes ν 8 ν 9 with frequencies ≈276 cm−1 the carbon cage oscillates with partial involvement of scandium atoms, the metal atoms oscillate symmetrically with respect to the center of the fullerene. These modes are IR inactive.
In contrast to Sc2@C60 (C3v), the pumpkin modes of Sc@C60 IR are active because of the involvement of the metal atom and the change in the position of the positive charge center during oscillation. The frequencies of the pumpkin modes of scandium mono-EMF are 231.91, 245.22, 254.84, 265.83, 269.56 cm−1, which are close to those of the di-EMF.
The vibrational mode ν 10 = 286.41 cm−1 is IR-active (see Figure 8) and represents shear vibrations of the Sc2 dimer along the Sc–Sc bond, with 7% involvement of metal atoms.
The vibrational mode ν 23 = 417.35 cm−1 is a combination of compression–stretching of the Sc2 dimer and vibrations of the carbon cage atoms with 4% of the metal atoms involved. This mode does not appear in the IR spectrum, since the vibrations of metal atoms are symmetric with respect to the fullerene center and therefore do not change the dipole moment of the molecule.
The vibrational mode ν 27 = 439.88 cm−1 is IR-active (see Figure 8); it is a combination of shear vibrations of the Sc2 dimer along the Sc–Sc bond and pentagon radial modes of the fullerene, with 1% involvement of metal atoms.
The vibrations with symmetric compression of the pentagonal rings of the carbon cage (pentagonal-pinch mode) have frequency ν 167 = 1509.15 cm−1 and are not IR-active. In an empty fullerene, this vibrational mode appears in the Raman spectrum and has a frequency of 1470 cm−1 [67].
In the spectrum of Sc2@C60 (C3v) in the frequency range up to 300 cm−1 there are 10 vibrational modes, of which 1 is IR-active (with a frequency of 286 cm−1) and 9 are IR-inactive. In the range from 300 to 400 cm−1, there are 11 modes, which correspond to 6 IR-active modes (with frequencies 340 (doublet of close lines), 367 (doublet), 398, 399 cm−1), 3 weak/low-intensity IR lines, and 2 IR-inactive modes. In the range 400–600 cm−1 there are 5 IR-active lines (with frequencies 440, 464, 465, 574, 574 cm−1), 3 weak IR lines, and 22 IR-inactive modes (30 modes in total). Of the 40 vibrational modes in the range 600–800 cm−1, there are 9 IR-active modes (602, 602, 622 (doublet), 654 (doublet), 657, 676, 677 cm−1), 7 weak IR lines, and 24 IR-inactive modes (see Figure 2).
In the spectral range 800–1000 cm−1, there are 12 Sc2@C60 (C3v) vibrational modes, of which 1 is weak IR mode and 11 are IR-inactive. In the range 1000–1200 cm−1, there are 16 vibrational modes, of which 4 are IR-active (with frequencies 1152, 1153, 1186 (doublet) cm−1), 2 are low-intensity IR lines, and 10 are IR-inactive. Of the 30 vibrational modes in the range 1200–1400 cm−1, there are 8 IR-active modes (1220 (doublet), 1226, 1342 (doublet), 1367, 1387 (doublet) cm−1), 3 low-intensity IR lines, and 19 IR-inactive modes. In the 1400–1600 cm−1 range, there are 10 IR-active lines (with the frequencies 1403, 1404 (doublet), 1452 (doublet), 1470, 1523 (doublet), 1527 (doublet) cm−1), 1 low-intensity IR line, and 20 IR-inactive modes (31 modes in total; see Figure 8).

2.6.2. Sc2@C60 ( C s )

The vibrational spectrum of Sc2@C60 with C s symmetry is presented in Figure 9.
The vibrational modes ν 1 = 78.95 cm−1 and ν 4 = 94.37 cm−1 represent shear vibrations of the Sc2 dimer along the axis perpendicular to the Sc–Sc bond, the fraction of metal atoms involvement being 10–17%. The vibrational modes are IR-active, which is due to the asymmetric position of metal atoms relative to the center of the fullerene.
The vibrational modes ν 2 = 79.24 cm−1 and ν 3 = 79.39 cm−1 are torsional vibrations of the Sc2 dimer around an axis passing perpendicularly through the middle of the metal–metal bond (lateral M–cage mode), the fraction of metal atoms involvement being 22–29%.
For Sc2@C60 ( C s ), the pumpkin modes ν 5 ν 9 have frequencies of 245.22 cm−1, 252.59 cm−1, 255.78 cm−1, 262.64 cm−1, and 268.56 cm−1. Of these vibrations, the highest IR activity has the ν 5 mode, for which the compression–stretching of the carbon cage is combined with shear vibrations of the Sc2 dimer along the Sc–Sc bond, the fraction of metal atoms involvement being 3%.
The vibrational mode ν 10 = 318.83 cm−1 is IR-active (see Figure 9) and represents shear vibrations of the Sc2 dimer along the Sc–Sc bond combined with deformations of the carbon cage, the fraction of metal atoms involvement being 2%.
The vibrational mode ν 13 = 347.14 cm−1 is a combination of compression–stretching of the Sc2 dimer and vibrations of the carbon cage atoms, with 2% involvement of metal atoms.
The vibrations at frequency ν 25 = 425.52 cm−1 are IR-active (see Figure 9). They are a combination of shear vibrations of the Sc2 dimer along the Sc–Sc bond and carbon cage radial mode, with 2% involvement of metal atoms.
In the spectrum of Sc2@C60 ( C s ), there are nine vibrational modes in the frequency range up to 300 cm−1 of which four are low-intensity IR modes and five are IR-inactive modes. In the range from 300 to 400 cm−1, there are 12 modes: 1 IR-active (319 cm−1), 4 low-intensity IR, and 7 IR-inactive. In the range 400–600 cm−1, there are 8 IR-active lines (with frequencies 403, 426, 456, 479, 505, 540, 562, 576 cm−1), 11 low-intensity IR lines, and 10 IR-inactive modes (29 modes in total). Additionally, 42 vibrational modes fall within the frequency range 600–800 cm−1, of which 7 are IR-active (604 (doublet), 630, 660, 671, 697, 706 cm−1), 12 low-intensity IR modes, and 23 IR-inactive modes (see Figure 2).
In the spectral range up to 800–1000 cm−1, there are 11 modes of Sc2@C60 ( C s ): 2 low-intensity IR and 9 IR-inactive modes. In the range 1000–1200 cm−1, there are 14 vibrational modes, of which 1 is IR-active (with a frequency of 1121 cm−1), 4 are low-intensity IR, and 9 are IR-inactive.
In the 1200–1400 cm−1 range, 34 vibrational modes were found: 14 IR-active (1203, 1205, 1217, 1223, 1258, 1275, 1350, 1354, 1358, 1377, 1382, 1386, 1389, 1396 cm−1), 9 low-intensity IR, and 11 IR-inactive.
Of the 29 modes falling within the range 1400–1600 cm−1, 10 lines are IR-active (with frequencies 1405, 1413, 1414 (doublet), 1448, 1455, 1468, 1483, 1509, 1512 cm−1), 9 low-intensity IR lines, and 10 IR-inactive modes (see Figure 9).

2.6.3. Y2@C60 ( C 3 v )

The vibrational spectrum of Y2@C60 with C 3 v symmetry is presented in Figure 10.
The vibrational modes ν 1 = 63.95 cm−1 and ν 2 = 65.99 cm−1 represent torsional vibrations of the Y2 dimer around the axis perpendicular to the Y–Y bond, the fraction of metal atoms involvement being 25%. This type of vibrations does not displace the positive charge center and does not change the dipole moment of the molecule, so it does not appear in IR spectra.
The vibrational modes ν 3 = 101.49 cm−1 and ν 4 = 102.91 cm−1 represent shear vibrations of the dimer Y2 along the axis passing perpendicularly through the middle of the Y–Y bond, the fraction of metal atoms involvement being 15%.
For the Y2@C60 di-EMF, yttrium atoms are involved only in the pumpkin mode ν 5 = 244.43 cm−1 (compression-extension of the Y2 dimer along the Y–Y bond), the fraction of metal atoms involvement is 5%. This mode is IR-inactive because the metal atoms oscillate symmetrically with respect to the fullerene center.
The vibrational mode ν 6 = 246.34 cm−1 is IR-active (see Figure 10) and represents shear vibrations of the Y2 dimer along the Y–Y bond (longitudinal M–cage mode, antisymmetric motion of metal atoms with respect to the fullerene center), with 10% involvement of metal atoms.
In pumpkin modes ν 7 ν 8 with frequencies 255–256 cm−1 the compression-extension of the carbon cage is perpendicular to the Y–Y bond, metal atoms are practically not involved, and these vibrations do not appear in the IR spectra, as well as in the empty fullerene C60. In pumpkin modes ν 9 ν 10 with frequencies 289 cm−1 vibrations of the carbon cage go with partial involvement of yttrium atoms, metal atoms oscillate symmetrically relative to the center of the fullerene. These modes are IR inactive.
In contrast to the corresponding Y2@C60 modes, the pumpkin modes of Y@C60 are IR-active. The non-symmetric involvement of the metal atom in the vibrations leads to a change in the dipole moment of the Y@C60 molecule, and therefore to appearance of the corresponding lines in the IR spectrum. The frequencies of the pumpkin modes of yttrium mono-EMF are 203.71, 236.67, 254.30, 262.87, 266.81 cm−1; these lines, except for the first one, are close in frequency to the di-EMF lines.
The vibrational mode ν 19 = 385.23 cm−1 is a combination of compression–stretching of the Y2 dimer and vibrations of the carbon cage atoms, the fraction of metal atoms involvement being 3%. This mode is not IR-active because the motion of metal atoms is symmetric with respect to the fullerene center.
The vibrational mode ν 26 = 431.05 cm−1 is IR-active (see Figure 10). It is a combination of shear vibrations of the Y2 dimer along the Y–Y bond and oscillating vibrations of the pentagonal rings of the fullerene, with 1% involvement of metal atoms.
The vibrations with symmetric compression of the pentagonal rings of the carbon cage (pentagonal-pinch mode) have frequency ν 167 = 1502.95 cm−1 and are not IR-active.
In the spectrum Y2@C60 ( C 3 v ) in the frequency range up to 300 cm−1 there are ten vibrational modes, of which one is IR-active (with a frequency of 246 cm−1) and nine are IR-inactive. From 300 to 400 cm−1 there are ten modes: five IR low-intensity lines and five IR-inactive modes. The range 400–600 cm−1 contains 5 IR-active lines (with frequencies 408, 409, 431, 573 (doublet) cm−1), 5 low-intensity IR lines, and 19 IR-inactive modes (29 modes in total). Forty-four oscillatory modes were found in the range 600–800 cm−1: 10 IR-active (610 (doublet), 628, 629, 635, 655, 657, 664, 680, 714 cm−1), 6 low-intensity IR lines, and 28 IR-inactive modes. (see Figure 3).
In the spectral range 800–1000 cm−1 there are ten Y2@C60 ( C 3 v ) modes, of which one is IR-active (938 cm−1) and nine are IR-inactive. In the range 1000–1200 cm−1, there are 18 vibrational modes, of which 2 are IR-active (with frequencies 1177, 1178 cm−1), 1 is weak IR line, and 15 modes are IR-inactive. Additionally, 35 vibrational modes fall within the frequency range 1200–1400 cm−1; of these, 13 are IR-active (1201, 1248, 1315, 1318, 1318, 1347 (doublet), 1364 (doublet), 1373 (doublet), 1384 (doublet) cm−1), 2 low-intensity IR lines and 20 modes are IR-inactive. In the 1400–1600 cm−1 range there are 7 IR-active lines (with frequencies 1429 (doublet), 1456, 1505, 1506, 1521, 1522 cm−1), 1 low-intensity IR line, and 16 IR-inactive modes (24 modes in total; see Figure 10).

2.7. Comparison of the Calculated Spectra

The spectra of mono- and di-EMFs of scandium are given in Figure 2 for comparison, the similar comparison for yttrium is given in Figure 3. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1 and ν 2 = 576 cm−1 [62]).

2.7.1. Spectra of Scandium Mono- and di-EMFs

In the range up to 300 cm−1 for mono-EMF scandium Sc@C60, one intense line at 55 cm−1 (lateral M-cage mode) is observed, while for di-EMF Sc2@C60 ( C 3 v ) there is also only one intense line at 286 cm−1 (pumpkin mode, which also includes the longitudinal M–cage mode). There are no intense IR lines of Sc2@C60 ( C s ) di-EMF in this spectral range. In the frequency range 300–400 cm−1 Sc@C60 has no intense lines. However, for Sc2@C60 ( C 3 v ), there is a group of lines with a frequency of 340–400 cm−1, and for Sc2@C60 ( C s ) an intense line with a frequency of 319 cm−1is also observed. In the 400–800 cm−1 domain, IR-active lines with close frequencies are observed for both mono-EMF and both isomers of di-EMF scandium.
In the 800–1000 cm−1 range, there are no IR-active lines of noticeable amplitude for scandium EMFs, Sc@C60, Sc2@C60 ( C 3 v ), Sc2@C60 ( C s ). In the range 1000–1110 cm−1 there are IR-active modes of the mono-EMF and no IR-active lines of noticeable amplitude for the di-EMFs. In the 1200–1600 cm−1 domain, the most intense spectral lines of scandium mono- and di-EMFs are observed. In general, due to its higher symmetry (compared to that of mono-EMF), the di-EMF has fewer IR-active lines in all parts of the spectrum. The spectra of mono- and di-EMFs differ most significantly in the region below 400 cm−1 (the spectral domain corresponding to vibrations involving the metal atoms), see Figure 2, Figure 6, Figure 8 and Figure 9.

2.7.2. Spectra of Yttrium Mono- and di-EMFs

In the range up to 400 cm−1, the Y@C60 mono-EMF shows one intense line at 53 cm−1 (lateral M–cage mode). For the di-EMF, Y2@C60 ( C 3 v ), there is also only one intense line in this range at 246 cm−1 (longitudinal M–cage mode). In the 400–600 cm−1 domain, most of the vibration frequencies of the mono- and di-EMF differ slightly, with the exception of the 500–550 cm−1 interval, in which there are IR-active modes of the mono-EMF while the di-EMF has no IR-active lines of noticeable amplitude.
In the 800–1000 cm−1 range, the di-EMF has an IR line at 938 cm−1, but and the mono-EMF has no IR-active lines of noticeable amplitude. In the range 1000–1150 cm−1, there are IR-active mono-EMF’s modes but no noticeable IR-active lines of the di-EMF. This is due to the reduced symmetry of the mono-EMF. The most intense spectral lines of the mono- and di-EMF are observed in the frequency range 1200–1600 cm−1.
In the spectra of yttrium mono- and di-EMFs, many of the vibrational modes are very close in frequency, since metal atoms are involved in these vibrations (for example, IR-active lines 1175 cm−1 for Y@C60 and 1177 cm−1 for Y2@C60).
In general, due to higher (compared to the mono-EMF) symmetry, the di-EMF spectrum has fewer IR-active lines in all parts of the spectrum, with the spectrum differing most significantly in the region below 400 cm−1 (the domain corresponding to the of vibrations that involve the metal atom), see Figure 3, Figure 7 and Figure 10.

2.7.3. Spectra of Sc and Y di-EMFs

The difference in the position of the most intense lines in the infrared spectra of M2@C60 from the IR-active lines of C60 is explained by a frequency shift due to the metal–carbon bond formation, as well as by the appearance of a number of IR-active lines that are symmetry-forbidden in the empty C60. In general, due to the lower symmetry of the Sc2@C60 ( C s ) molecule, its spectrum contains more IR-active lines compared to the spectra of M2@C60 with C 3 v symmetry.
In the range up to 300 cm−1 in the IR spectrum of Sc2@C60 ( C s ), there are no intense lines, for both EMFs with C 3 v symmetry; only one intense line is observed (longitudinal M–cage mode). Its frequency is 286 cm−1 for Sc2@C60 ( C 3 v ) and 246 cm−1 for Y2@C60 ( C 3 v ). The decrease in frequency can be explained by an increase in the atomic mass of the metal participating in the vibrations. In the 300–400 cm−1 spectral domain, a group of lines with frequencies 340–400 cm−1 is observed for Sc2@C60 ( C 3 v ). For Sc2@C60 ( C s ), a single intense line with a frequency of 319 cm−1 is observed, and there are no intense Y2@C60 ( C 3 v ) IR lines in this range.
In the frequency range 400–600 cm−1 for both EMFs with C 3 v symmetry, the most intense lines are very close in frequency: 440 cm−1and 574 cm−1for Sc2@C60 ( C 3 v ), 431 cm−1and 573 cm−1for Y2@C60 ( C 3 v ). In this range, the spectrum of Sc2@C60 ( C s ) also has close lines at 426 cm−1 and 576 cm−1, but the most intense are the 479 cm−1 and 505 cm−1 lines, which are absent in the spectra of M2@C60 with C 3 v symmetry. In the range 600–800 cm−1, a significant number of weak IR lines are observed for all M2@C60 under consideration. The strongest lines in this range are 622 cm−1 for Sc2@C60 ( C 3 v ), a doublet at 604 cm−1 for Sc2@C60 ( C s ), and 635 cm−1 for Y2@C60 ( C 3 v ). In the range from 800 to 1000 cm−1, there is an IR line at 938 cm−1of yttrium di-EMF only; no noticeable IR lines are found for both isomers of scandium di-EMF.
Near the frequency 1182 cm−1 of the IR-active line of empty C60 (marked with an arrow on the spectrum), an intense line is observed in the spectra of all considered M2@C60 di-EMFs. The metal atoms do not participate in these vibrations and have a weak effect on the parameters of this line.
The most intense lines of the IR spectra for all considered M2@C60 are located in the frequency range 1400–1500 cm−1. These lines correspond both to the IR-active line of empty C60 at 1429 cm−1 (which manifests in the IR spectra of EMFs), and to vibrational EMF modes that are IR-inactive for empty C60 fullerene; see Figure 8, Figure 9, Figure 10 and Figure 11.
Figure 11 compares the IR spectra below 800 cm−1 of the di-EMFs Sc2@C60 ( C s ), Sc2@C60 ( C 3 v ) and Y2@C60 ( C 3 v ). The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).

3. Methods

DFT calculation of the scandium subgroup EMFs were performed in Gaussian-09 [69] using the hybrid mPW3PBE function [70]. All the atoms were described by the all-electron valence jorge-DZP basis set [71]. The correspondence of the calculated structures to the potential energy surface minimum was checked by the absence of imaginary vibration frequencies.
To confirm the efficiency of the computational scheme, we perform some preliminary calculations of structural, spectral, and thermodynamic characteristics for C60 fullerene and metal carbides MC2 of the scandium subgroup (M = Sc, Y). For these carbide molecules, our calculations reproduce their isosceles triangle geometry known from experiment, the bond lengths being in good agreement with experimental data. These preliminary calculations allow us to evaluate the accuracy of the chosen quantum-chemical model of aromatic multi-atomic systems involving heavy atoms, to which the considered EMFs belong. Table 2 compares the results of our calculations with the experimental values of atomization enthalpies Δ H (at 0 K) and bond lengths [72,73,74,75], as well as the vibrational frequencies ν 1 ν 3 calculated in Refs. [76,77].
The next step in the verification of the computational scheme was to test it on the “empty” C60 fullerene calculations. Table 3 compares compares the results of our calculations with the experimental values standard enthalpy of formation, Δ H ([78], pp. 5–9), and bond lengths, r n m (length of the common edge of n- and m-membered rings in C60 fullerene) [79], as well as the vibrational frequencies ν 1 ν 4 measured in Ref. [62].
Because Sc and Y proton numbers are odd, the mono-EMFs M@C60 systems have open electronic shells, the same is the case for metal carbides, MC2. We used unrestricted Hartree–Fock (UHF) description of these systems. The ground electronic states of the mono-EMFs M@C60 are doublets (multiplicity M = 2 ), while the di-EMFs M2@C60 are characterized by singlet ground states ( M = 1 ).

4. Conclusions

Endohedral metallofullerenes (EMFs) are prospective materials with a number of possible applications in optoelectronics, spintronics, medicine and quantum computations. In the present work, we make DFT calculations of stable structures and IR spectra of Sc@C60 and Y@C60 monometallofullerenes (mono-EMFs), as well as those of Sc2@C60 and Y2@C60 dimetallofullerenes (di-EMFs). For the mono-EMFs, Sc@C60 and Y@C60, the lowest stable state corresponds to the isomer with the C s symmetry. The lowest stable states of di-EMFs, Sc2@C60 and Y2@C60, correspond to the C 3 v symmetry isomer. However, we detected another stable state of Sc2@C60 isomer with C s symmetry which has the energy by 0.11 eV greater than that of the C 3 v isomer.
Thus, the structures under our study were Sc@C60 ( C s ) , Y@C60 ( C s ) , Sc2@C60 ( C 3 v ) , Sc2@C60 ( C s ) and Y2@C60 ( C 3 v ) . We calculated the IR vibration spectra of these EMFs. For all the EMFs studied, the IR spectra consist of many lines, most of which do not appear in the IR spectra of empty C60 fullerene because the corresponding transitions are symmetry-forbidden.
For the spectra of mono- and di-EMFs of scandium and yttrium, we describe the metal–cage modes (i.e., the molecular vibrations involved the metal atoms). These are longitudinal and transverse modes, as well as rotational modes of the metal dimer. We also determine the directions of atomic motion for metal–cage modes and indicate the symmetry features of atomic displacements for both IR-active and IR-inactive vibrations. The corresponding illustrations and animations (short movies) are presented in the Supplementary Materials. Among the vibrations involving metal atoms, longitudinal vibrations along the metal–metal bond are characterized by the highest intensity in the IR spectrum of M2@C60 due to the displacement of the positive charge center.
We make a comparison between the vibrational modes of mono- and di-EMFs. Some IR-active modes of mono-EMFs become IR-inactive in di-EMFs. For example, the pumpkin vibrational mode in Sc@C60 and Y@C60, which is associated with simultaneous stretching-compression of all metal–carbon bonds, is IR-active. In the Sc2@C60 and Y2@C60 di-EMFs, similar modes are associated with symmetric compression–stretching of the metal dimer without displacement of the center of positive charge, so these oscillations are not IR-active.
We show that some Raman and “silent” modes of the empty C60 fullerene become IR-active in M@C60 and M2@C60 EMFs due to the influence of the encapsulated species.
In the EMF synthesis, the Sc2@C60 and Y2@C60 molecules have not been extracted experimentally to date, so the obtained IR spectra could help to detect and identify these EMFs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30163421/s1.

Author Contributions

Conceptualization, A.V.K.; methodology, I.V.N. and A.V.K.; software, I.V.N. and A.V.K.; investigation, I.V.N.; resources, G.E.K.; writing—original draft preparation, A.V.K.; writing—review and editing, V.E.C.; visualization, G.E.K.; project administration, V.E.C.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (grant number 24-22-00238).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Vibration frequencies ( ν k , cm−1) and line intensities ( I k , relative units.) for M@C60 endofullerenes.
Table A1. Vibration frequencies ( ν k , cm−1) and line intensities ( I k , relative units.) for M@C60 endofullerenes.
No. Sc@C 60 ( C s ) Y@C 60 ( C s )
ν k I k ν k I k
154.918810.503449.7620.5939
284.36821.306152.6225.1854
3231.90722.1295203.70652.1403
4245.22441.8977236.6692.3121
5254.84340.275254.29880.1129
6265.62960.036262.87070.0603
7269.56160.37266.8130.1857
8319.67951.2952304.11480.5722
9335.2551.2511331.64422.7561
10338.57880.9032340.89411.3791
11344.01581.3401342.15751.0795
12345.67430.0243344.25370.0151
13354.39391.5238351.51460.5949
14364.28190.1483363.03920.0128
15378.74770.8108365.91820.0229
16398.31750.78394.36821.7211
17399.3960.0095395.64820.4245
18399.7720.6038396.20641.315
19401.75260.2977398.85130.7967
20404.7190.3872400.74250.1307
21416.59782.3301404.14540.5953
22422.89691.1861418.66190.9549
23426.223311.2429421.885815.0228
24426.99512.4991424.76350.0654
25442.78123.82437.20063.3695
26463.27272.3697461.67010.6099
27479.04080.0182478.9560.2597
28484.88582.7619484.21661.3652
29493.45760.8953488.64311.863
30500.05510.0035492.25534.6305
31500.92491.8808493.83080.786
32506.46793.1632505.0552.8607
33511.68151.0673510.22211.2842
34519.321410.2078516.48766.8183
35523.89250.8604522.33630.2304
36529.554710.6873524.38465.0813
37538.04550.0683532.46120.0058
38543.03110.5084542.31370.3999
39552.39081.145548.57671.2618
40553.01140.0223552.34770.0395
41555.0420.3994554.09490.3178
42557.24582.8281554.6370.8385
43563.59562.3317564.75910.9204
44565.97370.1446564.93870.03
45571.96960.0044571.32550.02
46575.35724.6815572.71281.9636
47578.98912.1215578.487410.0132
48583.479315.3897578.53981.5721
49592.388926.0894591.381230.938
50606.278912.8187598.093619.172
51609.530.7846603.91772.4704
52637.21382.3095618.08119.1911
53640.91191.9561629.59121.8307
54651.60781.1549650.24020.6928
55662.3810.0207659.52930.256
56663.38351.9392661.66633.7551
57666.2410.1712664.57230.3112
58667.18080.3004665.29140.004
59668.7993.6108668.07544.4051
60670.30372.6129671.09781.0753
61674.18030.0278671.73882.1576
62676.67964.4683678.70480.0388
63677.09591.0939678.72113.099
64683.08841.2403683.92710.5006
65687.620911.1933686.179417.0234
66688.75710.8989689.22181.1416
67691.49074.1649690.61142.7627
68693.32451.6307692.87333.6523
69694.00591.4743695.28351.5051
70695.1732.2695695.98952.2944
71704.1010.0227701.91960.2835
72704.76250.494703.4391.4335
73705.05580.3317703.96150.2426
74705.06640.0904703.97660.3184
75709.0090.1873707.7470.011
76711.88090.087712.29430.3092
77722.28663.2397720.36892.4373
78727.75520.1452727.18940.031
79736.53691.9579736.17913.6256
80738.28880.1788750.92250.0559
81755.7610.0073756.58850.0462
82758.23150.0831759.39820.835
83760.55570.3114760.07180.0453
84762.37510.0989761.43140.0444
85778.5624.9404775.34876.5901
86783.86441.1869781.19522.1291
87788.02010.1788782.99620.4141
88788.63260.0309787.98930.0814
89799.84130.168793.98620.1776
90801.59830.2498799.19970.1369
91803.21580.239802.64870.2448
92803.24430.3458802.66510.0862
93825.7390.1767824.75250.1069
94826.80860.1544826.10070.0672
95829.5810.0003827.93330.0032
96941.46690938.91520.0013
97973.15061.3964970.17023.52
98975.19951.242973.3090.9045
99982.01831.0759978.99591.0072
100984.58740.6413981.63390.6234
1011001.07346.34851000.967910.0023
1021005.95380.67531004.65881.3841
1031010.11592.25041010.36932.1099
1041069.267819.85831050.239814.0723
1051104.89158.35981094.546418.5848
1061106.65633.64331097.60295.0212
1071121.6244.23921117.92923.8061
1081131.03342.62831128.21983.4907
1091134.32590.09411134.17280.55
1101144.18171.61871142.08742.9744
1111144.20333.68991143.11691.9164
1121144.8890.36291144.76540.7733
1131181.71759.49991175.11097.5938
1141203.92046.47621195.326410.1651
1151211.04412.79271203.16120.911
1161212.437311.77361204.180313.5954
1171224.96236.32761222.51036.9472
1181231.10920.46181228.91447.3897
1191233.97670.96761229.48573.0721
1201236.81937.99661234.88695.6621
1211240.119714.78991236.740218.5959
1221241.85461.18591238.99863.7997
1231252.35471.05431249.49363.4864
1241275.80723.39471266.15778.0192
1251285.43830.72191277.752.9424
1261293.47395.80091287.74588.2785
1271298.00865.6221296.52716.7523
1281306.05151.09481302.34942.4795
1291307.24290.69151305.75660.7031
1301309.60053.10421306.1044.7755
1311322.37850.00021321.06310.0612
1321331.21513.03441328.42622.5652
1331341.7770.27411336.09320.9261
1341345.5811.58011343.97781.7768
1351355.20060.34881347.25033.1169
1361356.73071.15971353.16070.6969
1371359.67250.19181357.45340.1664
1381362.42491.33461360.05861.8072
1391365.44160.14141362.54240.0419
1401365.95580.28961363.39122.7744
1411373.06133.53751367.87492.2178
1421373.81420.77221371.94789.1227
1431379.01991.52451375.0330.845
1441386.17310.01471379.10690.0336
1451388.75946.11981384.041926.5846
1461389.817920.85531385.421538.9756
1471396.46841.06231391.684768.3
1481398.887511.52021395.648911.8558
1491402.737822.3371396.413427.749
1501411.822274.35991399.810638.3463
1511433.88574.35621420.368879.1329
1521449.051128.24371435.827244.7941
1531456.410311.38051447.013195.5718
1541458.669368.37821455.203286.6558
1551467.192647.6121458.142122.6936
1561468.34451.72341468.210712.2601
1571474.434927.57061471.757337.0593
1581482.77828.86041479.74598.3834
1591484.697527.60661482.035823.1165
1601501.05016.36631495.60856.3091
1611508.5240.90381503.72951.0529
1621519.332627.11871520.533321.5776
1631539.64540.00091533.57795.6703
1641542.4815.67541534.4950.9343
1651544.59776.44491536.64776.2245
1661549.49580.49411541.31561.2692
1671551.69141.08171544.30374.0452
1681553.40860.40431545.60090.2033
1691563.07962.87731554.56611.1125
1701571.14122.47781561.57052.1147
1711572.85892.57821568.34562.6574
1721583.01244.01221575.15614.2037
1731598.03935.20151590.95578.5787
1741601.314514.11491595.629723.3172
1751604.69179.03111602.88638.3741
1761611.20098.63191605.77415.5851
1771612.3555.1461607.84938.5419
Table A2. Vibration frequencies ( ν k , cm−1) and line intensities ( I k , relative units.) for M2@C60 endofullerenes.
Table A2. Vibration frequencies ( ν k , cm−1) and line intensities ( I k , relative units.) for M2@C60 endofullerenes.
No. Sc 2 @C 60 ( C 3 v ) Sc 2 @C 60 ( C s ) Y 2 @C 60 ( C 3 v )
ν k I k ν k I k ν k I k
181.950.089178.952.503363.950.0000
282.68280.099179.240.005065.990.0000
389.0201079.390.5868101.490.0028
489.6563094.370.3545102.910.0038
5246.35210245.223.1424244.430.0001
6257.23190252.590.5416246.3414.3723
7257.64130255.780.2414254.990.0000
8276.09610.0002262.641.1927256.450.0000
9276.21720.0001268.560.0000288.630.0001
10286.406313.1518318.8311.8847288.830.0000
11332.28170.0017333.690.1506327.270.0005
12340.04854.6591338.530.1369334.060.6695
13340.25184.7108347.141.2174354.241.9465
14342.98381.2709355.530.0000354.802.0762
15366.87525.9916359.912.8679371.232.1276
16366.9716.0703369.471.0545371.302.1624
17386.9982.3818386.840.0332383.210.0017
18387.15142.4051390.960.0000383.340.0003
19391.58770.0028393.210.0315385.230.0000
20398.16624.4883395.120.5279388.180.0001
21398.50324.5393398.010.0000408.355.5303
22403.9120.0002403.069.8448408.565.5210
23417.35450421.450.8017416.580.0001
24430.38780425.300.1248425.180.0004
25430.76030425.5214.0096425.680.0000
26439.5180.0001431.304.1123431.0551.7812
27439.88277.4321446.190.0000451.870.0000
28440.16110.0555455.847.5080452.280.0001
29462.22930.0002478.680.0031460.010.6850
30464.25597.295479.2014.8063460.990.6246
31464.78757.2196486.831.0518466.970.0000
32477.03062.0317490.382.3426476.073.1430
33486.97040494.670.0000496.170.0000
34509.30850.0074505.3621.5666526.880.0260
35514.51943.2183506.190.9547529.700.9291
36514.7873.2154519.523.4110529.940.9939
37531.14890528.750.0927530.640.0078
38531.43580540.277.3260540.640.0000
39532.0680542.381.6787541.140.0000
40549.77590545.000.0000549.300.0000
41553.67640549.920.0000552.540.0001
42553.74940552.692.0966552.640.0000
43564.54950554.210.1566564.450.0002
44564.68920561.766.5229565.450.0000
45568.14530563.620.1104566.350.0003
46570.02650.0015567.020.0000566.440.0002
47573.75638.1059575.6011.4872573.1837.6595
48573.815938.1402577.371.2664573.2837.5027
49592.53560.0002590.650.0219586.860.0002
50592.72120.0001593.832.0420602.520.0014
51593.17750.0008601.530.0107602.900.0000
52602.23410.492603.6615.6539609.986.3385
53602.403111.3363603.6926.5221610.245.8881
54621.737421.346624.800.0000627.050.3042
55622.350619.5939630.136.0526628.038.6707
56623.97530.7065640.480.0340629.038.5735
57636.19442.2156643.480.4214634.9818.5265
58653.665211.7356647.900.0361655.084.8152
59654.10511.9753651.530.0000655.410.0195
60656.84440.0006659.554.5040655.950.0007
61657.37840.0061668.311.4273656.754.4998
62657.697311.8607670.181.1774663.975.3958
63669.62531.7236671.209.5921673.654.0864
64669.83021.6076671.530.0003673.814.2600
65676.29439.7942676.350.0295676.380.0002
66676.557110.223676.940.0000679.320.0718
67676.78210.0021677.700.2084679.654.2042
68677.30250.0013683.310.0944680.121.0623
69677.65510.0271684.950.0000680.395.4029
70678.29540.0001689.373.2574681.990.0059
71680.8140691.511.2116683.150.0003
72700.28680696.745.2250714.047.4142
73700.59390705.865.5792717.120.0017
74709.56480.0001706.511.2426717.520.0024
75710.30093.3806712.450.4658718.240.0002
76713.06453.5423716.040.0000725.541.3798
77713.46013.4636718.691.0675725.911.3945
78734.10410.094734.821.7653746.320.0250
79734.30330.1076737.290.0000746.410.0203
80739.35540742.021.3988748.030.0002
81739.4980743.590.0334755.770.0000
82753.17520.0003753.610.0246755.970.0000
83753.20070758.020.0011761.560.0931
84758.98850760.670.0000761.760.0957
85761.45120.0005763.460.0000765.271.6817
86761.51850.0006764.000.5207766.740.0008
87766.00230.1636764.660.0055767.100.0001
88775.8520774.780.0000778.280.0004
89776.01670786.861.0939785.520.0000
90791.4570786.910.5732792.360.0000
91791.62180791.720.1765792.470.0000
92800.05540799.880.0000794.110.0001
93805.66140804.420.1150794.860.0000
94805.85260805.470.2774802.170.0000
95808.0510805.490.0256802.240.0000
96825.67110824.750.0196824.020.0000
97825.77650824.970.0000824.130.0001
98831.54130830.380.0145826.710.0000
99940.9850.0001940.960.0000935.180.0055
100960.69193.0671962.804.0195938.137.5552
101970.06380.0169963.322.0807956.240.2956
102970.11740.015970.330.0000956.560.2961
103983.25680.0001982.160.0141981.950.0001
1041014.2511.06471011.960.81071001.440.0430
1051014.4871.07511013.890.02861001.550.0417
1061020.9970.39861017.280.17401015.300.0021
1071110.58801105.210.00001093.940.0000
1081110.62101105.310.10681094.020.0001
1091123.24401118.830.05501100.790.0000
1101123.40401120.574.91261100.900.0002
1111128.13501130.580.22431103.170.0000
1121151.7480.00021138.203.91361108.100.0970
1131152.3728.69741156.040.00001108.660.0907
1141152.5378.60671156.622.33231137.650.0003
1151163.41201157.630.10481155.210.0002
1161163.46801159.730.00001155.310.0001
1171165.101169.094.22201156.670.0000
1181186.29371.00261202.7011.41631177.4526.9316
1191186.43870.86421204.6317.93551177.5127.1322
1201213.3160.00121216.645.70631198.440.5152
1211219.5814.421223.068.60081198.570.3940
1221219.8214.49071223.332.03931201.288.2368
1231226.3077.08331223.390.00031206.870.0006
1241245.2950.00111238.880.42441225.290.0001
1251245.4210.00131250.034.04191225.480.0004
1261259.1810.00011253.011.73991243.670.6242
1271259.23701258.455.88221243.730.6234
1281262.8281.68481275.0112.22921247.575.8973
1291298.75901275.160.00001277.210.0002
1301298.82201294.820.44871277.340.0009
1311302.63401307.980.56971284.880.0011
1321318.45101313.540.51331305.020.0027
1331318.57701314.330.00001312.350.0058
1341333.3901324.821.91231312.680.0077
1351341.8218.41061337.410.00001315.44177.9984
1361341.87718.48871348.650.00001317.5423.9301
1371350.2601349.8317.68681317.9225.0840
1381361.420.00451354.196.73941331.820.0089
1391363.9601355.740.47371340.770.0082
1401363.97101358.226.12201346.100.0121
1411365.2890.00011364.800.04341346.7981.2490
1421367.39536.23361370.870.00101347.2382.6081
1431370.0491.82061374.980.00001349.600.0025
1441370.4131.94691376.5921.49971349.860.0208
1451376.2640.00231381.8212.22081364.2510.6934
1461386.94113.11411385.7725.69101364.4110.9053
1471387.03614.3251387.020.00461365.190.0058
1481390.8610.00021388.8548.52611365.280.0138
1491390.9930.00011392.170.07631371.400.0078
1501402.4140.00071395.8953.78251372.9322.0939
1511402.654185.5321398.460.00001373.1784.6566
1521403.807277.06291404.6515.01401384.2153.7793
1531403.904279.12881413.4235.32281384.3354.4713
1541417.7060.00511413.7385.35191393.550.0000
1551425.2420.00041414.49142.64771397.620.0007
1561425.2840.00091416.010.01361397.800.0010
1571451.926192.59611448.29108.86561428.93326.6849
1581452.03193.85821448.371.18581428.99325.9180
1591460.18301454.9732.79411446.090.0006
1601460.24601455.170.08501446.110.0003
1611462.1820.02921465.191.85841451.210.0085
1621470.15214.0831468.4745.29111455.8231.2813
1631485.1670.33961480.600.03671466.230.0005
1641485.2460.39571483.0519.28111466.460.0005
1651490.29901488.980.00001476.840.2428
1661490.4330.00011509.0916.20441476.980.2008
1671509.15301512.266.00891502.950.0001
1681523.0917.21131518.620.46191505.3016.6201
1691523.39717.45561532.340.34681505.5116.4199
1701527.18718.15741533.650.00001515.710.0000
1711527.33717.45341534.180.03261521.415.8033
1721529.740.00031538.520.00001521.605.8307
1731534.50801545.241.54451521.650.0559
1741542.9360.00141547.810.41481526.441.9181
1751548.9940.67981558.620.74721536.470.0000
1761550.79101559.330.36281536.720.0000
1771550.8901569.512.14751538.090.0005
1781573.82501576.790.00011559.160.0000
1791573.9701577.461.63811559.340.0000
1801578.16401592.760.63691576.680.0001

References

  1. Nie, H.; Zhao, C.; Shi, Z.; Jia, C.; Guo, X. Single-Molecule Fullerenes: Current Stage and Perspective. ACS Mater. Lett. 2022, 4, 1037–1052. [Google Scholar] [CrossRef]
  2. Shinohara, H. Endohedral metallofullerenes. Rep. Prog. Phys. 2000, 63, 843–892. [Google Scholar] [CrossRef]
  3. 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]
  4. Stevenson, S.; Fowler, P.W.; Heine, T.; Duchamp, J.C.; Rice, G.; Glass, T.; Harich, K.; Hajdu, E.; Bible, R.; Dorn, H.C. A stable non-classical metallofullerene family. Nature 2000, 408, 427–428. [Google Scholar] [CrossRef]
  5. Kako, M.; Nagase, S.; Akasaka, T. Functionalization of Endohedral Metallofullerenes with Reactive Silicon and Germanium Compounds. Molecules 2017, 22, 1179. [Google Scholar] [CrossRef]
  6. Jin, P.; Li, Y.; Magagula, S.; Chen, Z. Exohedral functionalization of endohedral metallofullerenes: Interplay between inside and outside. Coord. Chem. Rev. 2019, 388, 406–439. [Google Scholar] [CrossRef]
  7. Yamada, M.; Liu, M.T.H.; Nagase, S.; Akasaka, T. New Horizons in Chemical Functionalization of Endohedral Metallofullerenes. Molecules 2020, 25, 3626. [Google Scholar] [CrossRef] [PubMed]
  8. Dubrovin, V.; Avdoshenko, S.M. Conformational preferences of endohedral metallofullerenes on Ag, Au, and MgO surfaces: Theoretical studies. J. Comput. Chem. 2022, 43, 1614–1620. [Google Scholar] [CrossRef]
  9. Zhukov, V.V.; Erohin, S.V.; Churkin, V.D.; Vnukova, N.G.; Antipina, L.Y.; Elesina, V.I.; Visotin, M.A.; Tomashevich, Y.V.; Popov, M.Y.; Churilov, G.N.; et al. Feature of the Endohedral Metallofullerene Y@C82 and Gd@C82 Polymerization under High Pressure. J. Phys. Chem. C 2022, 126, 17366–17373. [Google Scholar] [CrossRef]
  10. Cioslowski, J. Electronic Structure Calculations on Endohedral Complexes of Fullerenes: Reminiscences and Prospects. Molecules 2023, 28, 1384. [Google Scholar] [CrossRef]
  11. Cai, W.; Zhang, M.; Echegoyen, L.; Lu, X. Recent advances in endohedral metallofullerenes. Fundam. Res. 2023, 5, 767–781. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, S.; Zhang, X.; Tan, X.; Li, H.; Dai, S.; Yao, B.; Liu, X.; He, Y.; Jin, F. Recent Progress on the Functionalization of Endohedral Metallofullerenes. Inorganics 2023, 11, 346. [Google Scholar] [CrossRef]
  13. Lu, X.; Feng, L.; Akasaka, T.; Nagase, S. Current status and future developments of endohedral metallofullerenes. Chem. Soc. Rev. 2012, 41, 7723–7760. [Google Scholar] [CrossRef]
  14. Yang, S.; Wang, X.; Chu, H.; Liu, W. Electrode-Supported Endohedral Metallofullerenes: Insights into the Confined Internal Dynamics. Inorg. Chem. 2024, 63, 6836–6844. [Google Scholar] [CrossRef] [PubMed]
  15. Amerikheirabadi, F.; Diaz, C.; Mohan, N.; Zope, R.R.; Baruah, T. A DFT analysis of the ground and charge-transfer excited states of Sc3N@Ih–C80 fullerene coupled with metal-free and zinc-phthalocyanine. Phys. Chem. Chem. Phys. 2018, 20, 25841–25848. [Google Scholar] [CrossRef]
  16. Trukhina, O.; Rudolf, M.; Bottari, G.; Akasaka, T.; Echegoyen, L.; Torres, T.; Guldi, D.M. Bidirectional Electron Transfer Capability in Phthalocyanine–Sc3N@Ih–C80 Complexes. J. Am. Chem. Soc. 2015, 137, 12914–12922. [Google Scholar] [CrossRef]
  17. Liang, Y.; Feng, L. Dual Electron Acceptor/Electron Donor Character of Endohedral Nitride Clusterfullerenes. Fullerenes Nanotub. Carbon Nanostruct. 2014, 22, 227–234. [Google Scholar] [CrossRef]
  18. Jones, L.O.; Mosquera, M.A.; Fu, B.; Schatz, G.C.; Ratner, M.A.; Marks, T.J. Germanium Fluoride Nanocages as Optically Transparent n-Type Materials and Their Endohedral Metallofullerene Derivatives. J. Am. Chem. Soc. 2019, 141, 1672–1684. [Google Scholar] [CrossRef] [PubMed]
  19. Okamura, N.; Yoshida, K.; Sakata, S.; Hirakawa, K. Electron transport in endohedral metallofullerene Ce@C82 single-molecule transistors. Appl. Phys. Lett. 2015, 106, 043108. [Google Scholar] [CrossRef]
  20. Xu, T.; Yin, H.; Yu, P.; He, Z.; Chen, N.; Shen, W.; Zhu, M.; Akasaka, T.; Lu, X. Ultraviolet Photodetectors Based on Dimetallofullerene Lu2@Cs(6)-C82 Nanorods. ACS Appl. Nano Mater. 2022, 5, 1683–1689. [Google Scholar] [CrossRef]
  21. Nam, J.S.; Seo, Y.; Han, J.; Lee, J.W.; Kim, K.; Rane, T.; Kim, H.D.; Jeon, I. Exploring Endohedral Metallofullerenes for Advanced Thin-Film Device Applications toward Next-Generation Electronics. Chem. Mater. 2023, 35, 8323–8337. [Google Scholar] [CrossRef]
  22. Sinitsa, A.S.; Chamberlain, T.W.; Zoberbier, T.; Lebedeva, I.V.; Popov, A.M.; Knizhnik, A.A.; McSweeney, R.L.; Biskupek, J.; Kaiser, U.; Khlobystov, A.N. Formation of Nickel Clusters Wrapped in Carbon Cages: Toward New Endohedral Metallofullerene Synthesis. Nano Lett. 2017, 17, 1082–1089. [Google Scholar] [CrossRef]
  23. Shui, Y.; Liu, D.; Zhao, P.; Zhao, X.; Ehara, M.; Lu, X.; Akasaka, T.; Yang, T. Element effects in endohedral metal–metal-bonding fullerenes M2@C82 (M = Sc, Y, La, Lu). J. Chem. Phys. 2023, 159, 244302. [Google Scholar] [CrossRef]
  24. Kandrashkin, Y.E.; Zaripov, R.B. Low-temperature motion of the scandium bimetal in endofullerene Sc2@C80(CH2Ph). Phys. Chem. Chem. Phys. 2023, 25, 31493–31499. [Google Scholar] [CrossRef]
  25. Wu, Y.; Zhou, Z.; Wang, Z. Stability and Electronic Properties of Mixed Rare-Earth Tri-Metallofullerenes YxDy3−x@C80 (x = 1 or 2). Molecules 2024, 29, 447. [Google Scholar] [CrossRef]
  26. Westerström, R.; Dreiser, J.; Piamonteze, C.; Muntwiler, M.; Weyeneth, S.; Brune, H.; Rusponi, S.; Nolting, F.; Popov, A.; Yang, S.; et al. An Endohedral Single-Molecule Magnet with Long Relaxation Times: DySc2N@C80. J. Am. Chem. Soc. 2012, 134, 9840–9843. [Google Scholar] [CrossRef]
  27. Hu, Z.; Yang, S. Endohedral metallofullerene molecular nanomagnets. Chem. Soc. Rev. 2024, 53, 2863–2897. [Google Scholar] [CrossRef]
  28. Wu, B.; Hu, J.; Cui, P.; Jiang, L.; Chen, Z.; Zhang, Q.; Wang, C.; Luo, Y. Visible-Light Photoexcited Electron Dynamics of Scandium Endohedral Metallofullerenes: The Cage Symmetry and Substituent Effects. J. Am. Chem. Soc. 2015, 137, 8769–8774. [Google Scholar] [CrossRef] [PubMed]
  29. Su, M.; Hu, Y.; Yang, S.; Yu, A.; Peng, P.; Yang, L.; Jin, P.; Su, B.; Li, F.F. Large-Area Endohedral Metallofullerene Single-Crystal Arrays for High-Performance Field-Effect Transistors and Photodetectors. Adv. Electron. Mater. 2022, 8, 2100753. [Google Scholar] [CrossRef]
  30. Chen, W.; Huang, M.; Wu, M.; Lei, Y. Endohedral Metallofullerenes: Unveiling Synthesis Mechanisms and Advancing Photoelectric Energy Conversion Applications. Top. Curr. Chem. 2025, 383, 14. [Google Scholar] [CrossRef] [PubMed]
  31. Martínez-Flores, C.; Basiuk, V.A. Ln@C60 endohedral fullerenes: A DFT analysis for the complete series from lanthanum to lutetium. Comput. Theor. Chem. 2022, 1217, 113878. [Google Scholar] [CrossRef]
  32. Putaud, T.; Chartrand, J.C.; Kalugina, Y.; Michaut, X.; Roy, P.N.; Ayotte, P. A simple confined rotor model to describe the ro-translational dynamics of water endofullerenes and to assign the ro-vibrational spectra of solid H2O@C60. J. Chem. Phys. 2025, 162, 144313. [Google Scholar] [CrossRef]
  33. Krisilov, A.; Nechaev, I.; Kotova, A.; Shikhaliev, K.; Chernov, V.; Zon, B. The role of the encapsulated atom in the vibrational spectra of La@C60–Lu@C60 lanthanide endofullerenes. Comput. Theor. Chem. 2015, 1054, 100–108. [Google Scholar] [CrossRef]
  34. Civiš, S.; Krisilov, A.V.; Ferus, M.; Nechaev, I.V.; Kubelík, P.; Chernov, V.E.; Zon, B.A. Vibrational spectra of La@C60 and Ce@C60 endohedral fullerenes: Influence of spin state multiplicity. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 254, 119593. [Google Scholar] [CrossRef] [PubMed]
  35. Nakagawa, A.; Nishino, M.; Niwa, H.; Ishino, K.; Wang, Z.; Omachi, H.; Furukawa, K.; Yamaguchi, T.; Kato, T.; Bandow, S.; et al. Crystalline functionalized endohedral C60 metallofullerides. Nat. Commun. 2018, 9, 3073. [Google Scholar] [CrossRef] [PubMed]
  36. Matsuo, Y.; Okada, H.; Ueno, H. Endohedral Lithium-Containing Fullerenes: Preparation, Derivatization, and Application; Springer: Singapore, 2017. [Google Scholar] [CrossRef]
  37. Ogawa, T.; Sugai, T.; Shinohara, H. Isolation and Characterization of Er@C60. J. Am. Chem. Soc. 2000, 122, 3538–3539. [Google Scholar] [CrossRef]
  38. Kanbara, T.; Kubozono, Y.; Takabayashi, Y.; Fujiki, S.; Iida, S.; Haruyama, Y.; Kashino, S.; Emura, S.; Akasaka, T. Dy@C60: Evidence for endohedral structure and electron transfer. Phys. Rev. B 2001, 64, 113403. [Google Scholar] [CrossRef]
  39. Kubozono, Y.; Maeda, H.; Takabayashi, Y.; Hiraoka, K.; Nakai, T.; Kashino, S.; Emura, S.; Ukita, S.; Sogabe, T. Extractions of Y@C60, Ba@C60, La@C60, Ce@C60, Pr@C60, Nd@C60, and Gd@C60 with Aniline. J. Am. Chem. Soc. 1996, 118, 6998–6999. [Google Scholar] [CrossRef]
  40. Wang, Z.; Omachi, H.; Shinohara, H. Non-Chromatographic Purification of Endohedral Metallofullerenes. Molecules 2017, 22, 718. [Google Scholar] [CrossRef]
  41. Wang, C.R.; Kai, T.; Tomiyama, T.; Yoshida, T.; Kobayashi, Y.; Nishibori, E.; Takata, M.; Sakata, M.; Shinohara, H. A Scandium Carbide Endohedral Metallofullerene: (Sc2C2)@C84. Angew. Chem. 2001, 40, 397–399. [Google Scholar] [CrossRef]
  42. Inoue, T.; Tomiyama, T.; Sugai, T.; Okazaki, T.; Suematsu, T.; Fujii, N.; Utsumi, H.; Nojima, K.; Shinohara, H. Trapping a C2 Radical in Endohedral Metallofullerenes: Synthesis and Structures of (Y2C2)@C82 (Isomers I, II, and III). J. Phys. Chem. B 2004, 108, 7573–7579. [Google Scholar] [CrossRef]
  43. Zheng, H.; Zhao, X.; Wang, W.W.; Yang, T.; Nagase, S. Sc2@C70 rather than Sc2C2@C68: Density functional theory characterization of metallofullerene Sc2@C70. J. Chem. Phys. 2012, 137, 014308. [Google Scholar] [CrossRef] [PubMed]
  44. Kurihara, H.; Lu, X.; Iiduka, Y.; Mizorogi, N.; Slanina, Z.; Tsuchiya, T.; Akasaka, T.; Nagase, S. Sc2C2@C80 Rather than Sc2@C82: Templated Formation of Unexpected C2v(5)−C80 and Temperature-Dependent Dynamic Motion of Internal Sc2C2 Cluster. J. Am. Chem. Soc. 2011, 133, 2382–2385. [Google Scholar] [CrossRef]
  45. Han, Y.; Zheng, H.; He, J.; Li, M.; Ehara, M.; Zhao, X. Stabilities Referring to the f Electron Effect in Erbium-Based Endohedral Metallofullerenes: Revisiting Er2@C80 and the Missing Er2C2@C78. Inorg. Chem. 2025, 64, 8302–8312. [Google Scholar] [CrossRef]
  46. Shen, Y.; Roselló, Y.; Abella, L.; Qiu, J.; Du, X.; Meng, Q.; Zheng, L.; Cao, Z.; He, Z.; Poblet, J.M.; et al. Fluoride Clusterfullerenes: Tuning Metal–Metal Bonding and Magnetic Properties via Single Fluorine Atom Doping. J. Am. Chem. Soc. 2024, 146, 34924–34933. [Google Scholar] [CrossRef]
  47. Yao, Y.R.; Shi, X.M.; Zheng, S.Y.; Chen, Z.C.; Xie, S.Y.; Huang, R.B.; Zheng, L.S. Atomically Precise Insights into Metal–Metal Bonds Using Comparable Endo-Units of Sc2 and Sc2C2. Chin. Chem. Soc. Chem. 2021, 3, 294–302. [Google Scholar] [CrossRef]
  48. Pan, C.; Shen, W.; Yang, L.; Bao, L.; Wei, Z.; Jin, P.; Fang, H.; Xie, Y.; Akasaka, T.; Lu, X. Crystallographic characterization of Y2C2n (2n = 82, 88–94): Direct Y–Y bonding and cage-dependent cluster evolution. Chem. Sci. 2019, 10, 4707–4713. [Google Scholar] [CrossRef]
  49. Inoue, T.; Kubozono, Y.; Kashino, S.; Takabayashi, Y.; Fujitaka, K.; Hida, M.; Inoue, M.; Kanbara, T.; Emura, S.; Uruga, T. Electronic structure of Eu@C60 studied by XANES and UV–VIS absorption spectra. Chem. Phys. Lett. 2000, 316, 381–386. [Google Scholar] [CrossRef]
  50. Yang, W.; Barbosa, M.F.d.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]
  51. Kwok, S. Complex Organics in Space: A Changing View of the Cosmos. Galaxies 2023, 11, 104. [Google Scholar] [CrossRef]
  52. El-Barbary, A. IR spectroscopic analysis of heterohedral metallofullerenes. Diam. Relat. Mater. 2023, 136, 110082. [Google Scholar] [CrossRef]
  53. Jiang, Y.; Li, Z.; Zhang, J.; Wang, Z. Endohedral Metallofullerene M2@C80: A New Class of Magnetic Superhalogen. J. Phys. Chem. C 2020, 124, 2131–2136. [Google Scholar] [CrossRef]
  54. Popov, A.A.; Avdoshenko, S.M.; Pendás, A.M.; Dunsch, L. Bonding between strongly repulsive metal atoms: An oxymoron made real in a confined space of endohedral metallofullerenes. Chem. Commun. 2012, 48, 8031–8050. [Google Scholar] [CrossRef]
  55. Zuo, T.; Xu, L.; Beavers, C.M.; Olmstead, M.M.; Fu, W.; Crawford, T.D.; Balch, A.L.; Dorn, H.C. M2@C79N (M = Y, Tb): Isolation and Characterization of Stable Endohedral Metallofullerenes Exhibiting M−M Bonding Interactions inside Aza[80]fullerene Cages. J. Am. Chem. Soc. 2008, 130, 12992–12997. [Google Scholar] [CrossRef]
  56. Slanina, Z.; Uhlík, F.; Lee, S.L.; Adamowicz, L.; Nagase, S. Lix@C60: Calculations of the Encapsulation Energetics and Thermodynamics. Int. J. Mol. Sci. 2008, 9, 1841–1850. [Google Scholar] [CrossRef]
  57. Hu, Y.H.; Ruckenstein, E. Endohedral Chemistry of C60-Based Fullerene Cages. J. Am. Chem. Soc. 2005, 127, 11277–11282. [Google Scholar] [CrossRef] [PubMed]
  58. Kobayashi, K.; Nagase, S. Structures and electronic states of endohedral dimetallofullerenes: M2@C80 (M = Sc, Y, La, Ce, Pr, Eu, Gd, Yb and Lu). Chem. Phys. Lett. 1996, 262, 227–232. [Google Scholar] [CrossRef]
  59. Kobayashi, K.; Nagase, S. Theoretical calculations of vibrational modes in endohedral metallofullerenes: La@C82 and Sc2@C84. Mol. Phys. 2003, 101, 249–254. [Google Scholar] [CrossRef]
  60. Inakuma, M.; Shinohara, H. Temperature-Dependent EPR Studies on Isolated Scandium Metallofullerenes: Sc@C82(I, II) and Sc@C84. J. Phys. Chem. B 2000, 104, 7595–7599. [Google Scholar] [CrossRef]
  61. Schettino, V.; Pagliai, M.; Ciabini, L.; Cardini, G. The Vibrational Spectrum of Fullerene C60. J. Phys. Chem. A 2001, 105, 11192–11196. [Google Scholar] [CrossRef]
  62. Bethune, D.S.; Meijer, G.; Tang, W.C.; Rosen, H.J.; Golden, W.G.; Seki, H.; Brown, C.A.; de Vries, M.S. Vibrational Raman and infrared spectra of chromatographically separated C60 and C70 fullerene clusters. Chem. Phys. Lett. 1991, 179, 181–186. [Google Scholar] [CrossRef]
  63. Krause, M.; Popov, V.N.; Inakuma, M.; Tagmatarchis, N.; Shinohara, H.; Georgi, P.; Dunsch, L.; Kuzmany, H. Multipole induced splitting of metal-cage vibrations in crystalline endohedral D2dM2@C84 dimetallofullerenes. J. Chem. Phys. 2004, 120, 1873–1880. [Google Scholar] [CrossRef]
  64. Schettino, V.; Remigio Salvi, P.; Bini, R.; Cardini, G. On the vibrational assignment of fullerene C60. J. Chem. Phys. 1994, 101, 11079–11081. [Google Scholar] [CrossRef]
  65. Iglesias-Groth, S.; Cataldo, F.; Manchado, A. Infrared spectroscopy and integrated molar absorptivity of C60 and C60 fullerenes at extreme temperatures. Mon. Not. R. Astron. Soc. 2011, 413, 213–222. [Google Scholar] [CrossRef]
  66. Tamuliene, J. Electronic and Vibrational Spectra of C60 and Its Ions. Fuller. Nanotub. Carbon Nanostructures 2015, 23, 187–195. [Google Scholar] [CrossRef]
  67. Menéndez, J.; Page, J.B. Vibrational spectroscopy of C60. In Light Scattering in Solids VIII: Fullerenes, Semiconductor Surfaces, Coherent Phonons; Cardona, M., Güntherodt, G., Eds.; Springer: Berlin/Heidelberg, Germany, 2000; pp. 27–95. [Google Scholar] [CrossRef]
  68. Tuchin, A.V. A realignment and activation of vibrational modes of the fullerene C60 and C70 in the electric field. Condens. Matter Interphases 2014, 16, 323–336. [Google Scholar]
  69. 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 Revision D.1; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  70. Adamo, C.; Barone, V. Exchange functionals with improved long-range behavior and adiabatic connection methods without adjustable parameters: The mPW and mPW1PW models. J. Chem. Phys. 1998, 108, 664–675. [Google Scholar] [CrossRef]
  71. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef]
  72. Chandrasekharaiah, M.; Gingerich, K. Chapter 86 Thermodynamic properties of gaseous species. In Handbook on the Physics and Chemistry of Rare Earths; Karl, A., Gschneidner, J., Eyring, L., Eds.; Elsevier: Amsterdam, The Netherlands, 1989; Volume 12, pp. 409–431. [Google Scholar] [CrossRef]
  73. Kohl, F.J.; Stearns, C.A. Vaporization Thermodynamics of Yttrium Dicarbide–Carbon System and Dissociation Energy of Yttrium Dicarbide and Tetracarbide. J. Chem. Phys. 1970, 52, 6310–6315. [Google Scholar] [CrossRef]
  74. Min, J.; Halfen, D.; Ziurys, L. Fourier transform microwave/millimeter-wave spectroscopy of the ScC2 ( X ˜ 2A1) radical: A model system for endohedral metallofullerenes. Chem. Phys. Lett. 2014, 609, 70–75. [Google Scholar] [CrossRef]
  75. Halfen, D.; Min, J.; Ziurys, L. The Fourier transform microwave spectrum of YC2 ( X ˜ 2A1) and its 13C isotopologues: Chemical insight into metal dicarbides. Chem. Phys. Lett. 2013, 555, 31–37. [Google Scholar] [CrossRef]
  76. Burton, M.A.; Cheng, Q.; Halfen, D.T.; Lane, J.H.; DeYonker, N.J.; Ziurys, L.M. The structure of ScC2 ( X ˜ 2A1): A combined Fourier transform microwave/millimeter-wave spectroscopic and computational study. J. Chem. Phys. 2020, 153, 034304. [Google Scholar] [CrossRef] [PubMed]
  77. Zhao, Y.; Wang, M.; Yang, C.; Ma, X.; Li, J. DFT calculations for anharmonic force field and spectroscopic constants of YC2 and its 13C isotopologues. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 191, 382–388. [Google Scholar] [CrossRef] [PubMed]
  78. Standard Thermodynamic Properties of Chemical Substances. In CRC Handbook of Chemistry and Physics, 97th ed.; Haynes, W.M., Lide, D.R., Bruno, T.J., Eds.; CRC Press/Taylor and Francis: Boca Raton, FL, USA, 2017; pp. 5–9. [Google Scholar]
  79. Hedberg, K.; Hedberg, L.; Bethune, D.S.; Brown, C.A.; Dorn, H.C.; Johnson, R.D.; Vries, M.D. Bond Lengths in Free Molecules of Buckminsterfullerene, C60 from Gas-Phase Electron Diffraction. Science 1991, 254, 410–412. [Google Scholar] [CrossRef]
Figure 1. Structure of M2@C60 endofullerenes: (a) M@C60 structure, numeration of carbon atoms closest to the metal atom. (b) Sc2@C60, isomer of C3v symmetry. (c) Sc2@C60, isomer of Cs symmetry. (d) Y2@C60, isomer of C3v symmetry.
Figure 1. Structure of M2@C60 endofullerenes: (a) M@C60 structure, numeration of carbon atoms closest to the metal atom. (b) Sc2@C60, isomer of C3v symmetry. (c) Sc2@C60, isomer of Cs symmetry. (d) Y2@C60, isomer of C3v symmetry.
Molecules 30 03421 g001
Figure 2. Comparison of IR spectra of scandium mono- and di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Figure 2. Comparison of IR spectra of scandium mono- and di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Molecules 30 03421 g002
Figure 3. Comparison of IR spectra of yttrium mono- and di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Figure 3. Comparison of IR spectra of yttrium mono- and di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Molecules 30 03421 g003
Figure 4. IR-active vibrational modes in M2@C60 di-EMFs (the arrows show the directions of atomic displacements): (a) ν 6 M–cage mode in Y 2 @ C 60 ( C 3 v ) . (b) ν 16 “silent” mode in Y 2 @ C 60 ( C 3 v ) (counter-torsional vibrations of the cage hemispheres). (c) ν 26 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). (d) ν 28 “breathing” radial mode in Sc 2 @ C 60 ( C s ) . (e) ν 48 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). (f) ν 57 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). A visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).
Figure 4. IR-active vibrational modes in M2@C60 di-EMFs (the arrows show the directions of atomic displacements): (a) ν 6 M–cage mode in Y 2 @ C 60 ( C 3 v ) . (b) ν 16 “silent” mode in Y 2 @ C 60 ( C 3 v ) (counter-torsional vibrations of the cage hemispheres). (c) ν 26 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). (d) ν 28 “breathing” radial mode in Sc 2 @ C 60 ( C s ) . (e) ν 48 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). (f) ν 57 radial mode in Y 2 @ C 60 ( C 3 v ) (strong line in IR spectrum). A visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).
Molecules 30 03421 g004
Figure 5. IR-active (unless otherwise specified) vibrational modes in M@C60 mono-EMFs (the arrows show the directions of atomic displacements): (a) ν 1 M–cage mode in Sc @ C 60 ( C s ) . (b) ν 3 M–cage “pumpkin” mode in Sc @ C 60 ( C s ) . (c) ν 6 “pumpkin” (IR-inactive) mode in Sc @ C 60 ( C s ) . (d) ν 48 radial mode in Sc @ C 60 ( C s ) (strong line in IR spectrum). A visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).
Figure 5. IR-active (unless otherwise specified) vibrational modes in M@C60 mono-EMFs (the arrows show the directions of atomic displacements): (a) ν 1 M–cage mode in Sc @ C 60 ( C s ) . (b) ν 3 M–cage “pumpkin” mode in Sc @ C 60 ( C s ) . (c) ν 6 “pumpkin” (IR-inactive) mode in Sc @ C 60 ( C s ) . (d) ν 48 radial mode in Sc @ C 60 ( C s ) (strong line in IR spectrum). A visualization of the atom motion for some vibrational modes is presented in the Supplementary Materials (short movies).
Molecules 30 03421 g005
Figure 6. IR spectrum of Sc@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Figure 6. IR spectrum of Sc@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Molecules 30 03421 g006
Figure 7. IR spectrum of Y@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Figure 7. IR spectrum of Y@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Molecules 30 03421 g007
Figure 8. Vibrational spectrum of Sc2@C60 endofullerene (isomer of C 3 v symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Figure 8. Vibrational spectrum of Sc2@C60 endofullerene (isomer of C 3 v symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Molecules 30 03421 g008
Figure 9. Vibrational spectrum of Sc2@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Figure 9. Vibrational spectrum of Sc2@C60 endofullerene (isomer of C s symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Molecules 30 03421 g009
Figure 10. Vibrational spectrum of Y2@C60 endofullerene (isomer of C 3 v symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Figure 10. Vibrational spectrum of Y2@C60 endofullerene (isomer of C 3 v symmetry). The black arrows show the positions of four IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1, ν 2 = 576 cm−1, ν 3 = 1182 cm−1 and ν 4 = 1429 cm−1 [62]).
Molecules 30 03421 g010
Figure 11. Comparison of IR spectra of Sc2@C60 ( C s ), Sc2@C60 ( C 3 v ) and Y2@C60 ( C 3 v ) di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Figure 11. Comparison of IR spectra of Sc2@C60 ( C s ), Sc2@C60 ( C 3 v ) and Y2@C60 ( C 3 v ) di-EMFs. The black arrows show the positions of two IR lines of the empty C60 fullerene ( ν 1 = 527 cm−1and ν 2 = 576 cm−1 [62]).
Molecules 30 03421 g011
Table 1. Interatomic distances in the calculated EMF structures.
Table 1. Interatomic distances in the calculated EMF structures.
Sc@C 60 ( C s ) Y@C 60 ( C s ) Sc 2 @C 60 ( C 3 v ) Sc 2 @C 60 ( C s ) Y 2 @C 60 ( C 3 v )
r(M–C1), Å2.172.342.262.242.36
r(M–C2), Å2.242.402.262.242.36
r(M–C3), Å2.332.482.262.252.36
r(M–M), Å3.253.113.13
Table 2. The calculated (this work) and experimental characteristics of metal carbides.
Table 2. The calculated (this work) and experimental characteristics of metal carbides.
ScC 2 YC 2
calc. expt. calc. expt.
Δ H , kJ/mol1183118212391254
r(M–C), Å2.0462.0572.1912.187
r(C–C), Å1.2781.2591.2771.270
ν 1 , cm−1335330.0337342.0
ν 2 , cm−1662634.6604581.1
ν 3 , cm−118161767.918261839.7
Table 3. The calculated (this work) and experimental characteristics of the “empty” C60 fullerene.
Table 3. The calculated (this work) and experimental characteristics of the “empty” C60 fullerene.
calc.expt.diff.
Δ H , kJ/mol240525024.2%
r 5 6 , Å1.4511.4520.1%
r 6 6 , Å1.3971.397<0.1%
ν 1 , cm−15235270.76%
ν 2 , cm−15955763.2%
ν 3 , cm−1123311824.1%
ν 4 , cm−1149714294.6%
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

Krisilov, A.V.; Nechaev, I.V.; Chernov, V.E.; Kallu, G.E. DFT Calculations of Structure and IR Spectra of M@C60 and M2@C60 Endofullerenes (M=Sc and Y). Molecules 2025, 30, 3421. https://doi.org/10.3390/molecules30163421

AMA Style

Krisilov AV, Nechaev IV, Chernov VE, Kallu GE. DFT Calculations of Structure and IR Spectra of M@C60 and M2@C60 Endofullerenes (M=Sc and Y). Molecules. 2025; 30(16):3421. https://doi.org/10.3390/molecules30163421

Chicago/Turabian Style

Krisilov, Alexey V., Igor V. Nechaev, Vladislav E. Chernov, and Gie Eli Kallu. 2025. "DFT Calculations of Structure and IR Spectra of M@C60 and M2@C60 Endofullerenes (M=Sc and Y)" Molecules 30, no. 16: 3421. https://doi.org/10.3390/molecules30163421

APA Style

Krisilov, A. V., Nechaev, I. V., Chernov, V. E., & Kallu, G. E. (2025). DFT Calculations of Structure and IR Spectra of M@C60 and M2@C60 Endofullerenes (M=Sc and Y). Molecules, 30(16), 3421. https://doi.org/10.3390/molecules30163421

Article Metrics

Back to TopTop