Next Article in Journal
Exploring Bacterial Cellulose and a Biosurfactant as Eco-Friendly Strategies for Addressing Pharmaceutical Contaminants
Previous Article in Journal
Chitosan Biocomposites with Variable Cross-Linking and Copper-Doping for Enhanced Phosphate Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability and Electronic Properties of Mixed Rare-Earth Tri-Metallofullerenes YxDy3-x@C80 (x = 1 or 2)

1
Key Laboratory of Advanced Light Conversion Materials and Biophotonics, Department of Chemistry, Renmin University of China, Beijing 100872, China
2
School of Materials Science and Engineering, Dalian Jiaotong University, Dalian 116028, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(2), 447; https://doi.org/10.3390/molecules29020447
Submission received: 25 December 2023 / Revised: 11 January 2024 / Accepted: 13 January 2024 / Published: 16 January 2024
(This article belongs to the Special Issue Theoretical Research of Carbon Nanomaterials)

Abstract

:
Tri-metallofullerenes, specifically M 3 @ C 80 where M denotes rare-earth metal elements, are molecules that possess intriguing magnetic properties. Typically, only one metal element is involved in a given tri-metallofullerene molecule. However, mixed tri-metallofullerenes, denoted as M1xM23-x@C80 (x = 1 or 2, M1 and M2 denote different metal elements), have not been previously discovered. The investigation of such mixed tri-metallofullerenes is of interest due to the potential introduction of distinct properties resulting from the interaction between different metal atoms. This paper presents the preparation and theoretical analysis of mixed rare-earth tri-metallofullerenes, specifically YxDy3−x@C80 (x = 1 or 2). Through chemical oxidation of the arc-discharge produced soot, the formation of tri-metallofullerene cations, namely Y 2 D y @ C 80 + and Y D y 2 @ C 80 + , has been observed. Density functional theory (DFT) calculations have revealed that the tri-metallofullerenes YxDy3−x@C80 (x = 1 or 2) exhibit a low oxidation potential, significantly lower than other fullerenes such as C60 and C70. This low oxidation potential can be attributed to the relatively high energy level of a singly occupied orbital. Additionally, the oxidized species demonstrate a large HOMO-LUMO gap similar to that of YxDy3−xN@C80, underscoring their high chemical stability. Theoretical investigations have uncovered the presence of a three-center two-electron metal–metal bond at the center of Y 2 D Y @ C 80 + and Y D y 2 @ C 80 + . This unique multi-center bond assists in alleviating the electrostatic repulsion between the metal ions, thereby contributing to the overall stability of the cations. These mixed rare-earth tri-metallofullerenes hold promise as potential candidates for single-molecule magnets.

1. Introduction

Fullerenes, which are carbon allotropes characterized by their cage-like structure, exhibit a notable structural feature of possessing a cavity with dimensions smaller than a nanometer. This inherent property allows for the accommodation of metal atoms within the internal space of the fullerene cages, leading to the creation of metallofullerenes. The synthesis, structures, and properties of metallofullerene molecules have undergone thorough examination [1,2,3,4,5,6,7]. The incorporation of metal atoms or clusters enhances the optical, electrical, and magnetic properties of fullerene molecules, offering promising prospects for development in nonlinear optics, single-molecule magnets, fluorescent materials, and various other fields [8,9,10,11,12]. In recent years, researchers have continuously expanded the types and forms of fullerenes by exploring new structures of metallofullerenes, which provides important inspiration for the development of novel molecular-based functional materials.
The internal cavity of a fullerene molecule typically falls within the sub-nanometer scale, capable of accommodating one, two, three, or more atoms [13,14,15,16,17,18,19]. M @ C 82 stands out as the most thoroughly examined mono-metallofullerene [20,21,22,23,24,25,26], primarily attributed to its remarkable production yield. The fullerene structure of M @ C 82 can undergo functionalization, leading to the creation of various derivatives of metallofullerenes. An intriguing cationic mono-metallofullerene is L i @ C 60 + . When compared to C60, L i @ C 60 + demonstrates enhanced hydrogenation reactivity [27]. The most-extensively studied di-metallofullerenes are M 2 @ C 80 [28,29,30]. Theoretical calculations for M 2 @ C 80 indicate that lanthanide element dimers like L a 2 and C e 2 transfer six electrons to the carbon cage, resulting in the electron configuration of M 2 6 + @ C 80 6 . As the metal atom has contributed all its valence electrons, there is no metal–metal bond in M 2 @ C 80 (M = La, Ce). On the other hand, lanthanide metal dimers, such as G d 2 , T b 2 , D y 2 , and E r 2 , contribute five electrons to the carbon cage, forming the electron configuration of M 2 5 + @ C 80 5 , while retaining a metal bond occupied by a single electron. The neutral form of M 2 @ C 80 (M = Gd, Tb, Dy, Er) is unstable due to the open-shell electronic structure of the outer carbon cage. However, covalent derivatization can stabilize these di-metallofullerenes. Of particular interest are the derivatives of T b 2 @ C 80 and D y 2 @ C 80 , which exhibit unique single-molecule magnet behavior [31].
In the case of metallofullerenes encapsulating three or more metal atoms, it is common to include one or more non-metal atoms within the fullerene cage. Examples include S c 3 N @ C 80 , V S c 2 N @ C 80 , S c 3 C 2 @ C 80 , S c 4 C 2 @ C 80 , T i 3 C 3 @ C 80 , and D y 3 C 2 @ C 80 , etc. [19,32,33,34,35,36]. This is attributed to the significant Coulomb repulsion between the metal ions, which can be alleviated by the presence of non-metal atoms. Besides these cluster fullerenes, several reports have been published on tri-metallofullerenes without non-metal mediators, including E r 3 @ C 74 , Y 3 @ C 80 , S m 3 @ C 80 , T m 3 @ C 80 , and others [37,38,39,40,41,42]. Popov et al. have postulated the presence of a pseudo atom at the core of the Y 3 @ C 80 molecule, simulating the N atom in Y 3 N @ C 80 [41]. The interaction between the Y atoms and the pseudo atom mirrors that between the Y atoms and the N atom in Y 3 N @ C 80 . Notably, the structure of S m 3 @ C 80 has been elucidated using single-crystal X-ray diffraction among these tri-metallofullerenes [42]. Recently, our group reported the successful extraction of T m 3 @ C 80 from arc-discharge-produced soot through chemical oxidation [43]. Theoretical investigations have suggested the presence of a three-center two-electron metal–metal bonding in these tri-metallofullerenes. Moreover, larger tri-metallofullerenes have been identified in the gas phase via laser ablation [44,45,46,47].
To date, all the previously reported tri-metallofullerenes have exclusively utilized a single category of metal elements. The feasibility of encapsulating diverse metal atoms within the fullerene cage to generate tri-metallofullerenes with mixed elements remains uncertain. This study endeavors to investigate the synthesis of mixed rare-earth tri-metallofullerenes, specifically YxDy3−x@C80 (x = 1 or 2). DFT calculations have demonstrated their effectiveness in examining the structures and properties of metallofullerenes, establishing them as a reliable and robust method [10]. In this work, we assess the stability and electronic characteristics of YxDy3−x@C80 (x = 1 or 2) through DFT calculations.

2. Results and Discussion

Following the arc-discharge process using a Y and Dy precursor mixture (with a molar ratio of Y:Dy = 2:1), we conducted the chemical oxidation of the resulting soot. Figure 1 displays the mass spectrum of the oxidized products. The most prominent peak corresponds to Y 3 @ C 80 . Signals indicative of mixed rare-earth tri-metallofullerenes, specifically Y 2 D y @ C 80 and Y D y 2 @ C 80 , were also observed. This marks the first observation of the existence of mixed rare-earth tri-metallofullerenes. In this work, the mass spectra were measured in the positive mode. Typically, the peak intensity in the mass spectra may not accurately represent the yields of different metallofullerenes due to their varying ionization energies. However, in this study, the trimetallofullerenes ( Y 3 @ C 80 , Y2Dy@C80 and YDy2@C80) have very similar ionization energies, with calculations yielding values of 5.15, 5.14, and 5.13 eV, respectively. Consequently, based on the intensity of the mass peaks, the yields decrease in the following order: Y 3 @ C 80 > Y2Dy@C80 > YDy2@C80. Other peaks in the mass spectrum correspond to mono-metallofullerens M2@C2n and di-metallofullerenes M 2 @ C 2 n (M = Y or Dy). Fullerenes C60 and C70 were also produced in the arc-discharge process. However, their solubility in dichloromethane—the solvent used—is low. On the other hand, oxidized metallofullerenes readily dissolve in dichloromethane. As a result, the metallofullerenes were concentrated in the extract.
Previous investigations have suggested that both Y 3 @ C 80 and T m 3 @ C 80 share the same fullerene cage as Y 3 N @ C 80 and T m 3 N @ C 80 , specifically the I h -symmetric C 80 cage [41,43]. It should be noted that the symmetry of this fullerene is lowered to C 3 V due to the Jahn–Teller symmetry reduction. The cage obtains its highest symmetry only when filled with metals [48]. To ascertain whether Y2Dy@C80 and YDy2@C80 also adopt this particular fullerene cage, we conducted DFT calculations. Prior to this study, Popov et al. performed theoretical calculations to establish the energy order of C 80 6 isomers [49]. We considered the nine most stable C 80 6 isomers as potential cages for encapsulating the Y and Dy atoms, forming Y2Dy@C80 and YDy2@C80. Consequently, we obtained nine isomers for Y2Dy@C80 and YDy2@C80. The relative energies, computed using the PBE0 functional, are presented in Table 1 and Table 2. Our analysis reveals that the Ih-symmetric cage corresponds to the lowest energy isomer for both Y2Dy@C80 and YDy2@C80. Their molecular structures are depicted in Figure 2.
When encapsulating multiple metal atoms within a fullerene cage, it is often necessary to introduce one or two non-metal atoms between them to counteract the strong Coulomb repulsion. Examples of this phenomenon include S c 3 C 2 @ C 80 , D y 3 C 2 @ C 80 , and other cases [33,36,50]. Theoretical confirmation exists showing that certain metal carbide cluster fullerenes, such as MxCy@C2n, possess lower energy compared to their metallofullerene counterparts Mx@C2n+y. To assess the stability of Y2Dy@Ih−C80 (referred to as Y2Dy@C80 hereafter) and Y2DyC2@C78, we also conducted energy calculations for Y2DyC2@C78. The computational results indicate that Y2DyC2@C78 exhibits higher energy than Y2Dy@C80, as detailed in Table 3. Consequently, it can be concluded that Y2Dy@C80 is thermodynamically more stable than its metallic carbon cluster counterpart, Y2DyC2@C78. A similar conclusion was drawn for YDy2@C80 (Table 4).
The stability of Y2Dy@C80 and Y D y 2 @ C 80 was assessed by computing the binding energies ( Δ E b ) between the metal cluster and fullerene cage using hypothetical reactions (1) and (2). The definition of binding energy is as follows: Δ E b = E(fullerene) + E(metal cluster) − E(metallofullerene). Higher binding energy values indicate a greater stability of metallofullerenes. The binding energy calculated for Y2Dy2@C80 is 281 kcal/mol. On the other hand, the binding energy calculated for YDy2@C80 is smaller, specifically 273 kcal/mol. The computed binding energy for Y3@C80 is 289 kcal/mol. This sequence of the binding energy for Y3@C80, Y2Dy@C80, and YDy2@C80 aligns with the order of their peak intensities in the mass spectrum. Consequently, it is reasonable to infer that the yield of tri-metallofullerenes is correlated with the binding energy. We have computed the van der Waals volume, which refers to the space enclosed within the isosurface with an electron density of 0.001 a.u., for C80, Y3@C80, Y2Dy@C80, and YDy2@C80. The obtained values are 820, 870, 875, and 876 Å3 for C80, Y3@C80, Y2Dy@C80, and YDy2@C80, respectively. It is evident that the volume expands due to the incorporation of the metal cluster. The ionic radius of Dy3+ (0.091 nm) is larger than that of Y3+ (0.089 nm). According to the computed binding energies, it can be inferred that the Y2Dy cluster is better suited than the YDy2 cluster for encapsulation within the C80 cage.
For comparison, we also calculated the binding energy for Y2DyN@C80 and YDy2N@C80, both of which are nitride clusterfullerenes known for their exceptional stability (reactions (3) and (4)). Previous studies have examined the stability of the nitride clusterfullerene M3N@C80 for different metal atoms. These studies have revealed that certain rare-earth elements, such as La and Nd, are too large to fit within the Ih-C80 fullerene cage, while Dy and Y atoms possess suitable sizes. The computed binding energies for La3@C80 and Nd3@C80 are 217 and 237 kcal/mol, respectively, which are significantly lower than those of Y3@C80, Y2Dy@C80, and YDy2@C80. Signals corresponding to La3@C80 and Nd3@C80 were not detected in the oxidized products. We attribute this observation to the substantial size of the La and Nd atoms. Our calculations demonstrate that the binding energies of Y2DyN@C80 and YDy2N@C80 are 269 and 257 kcal/mol, respectively. The binding energies of Y2Dy@C80 and YDy2@C80 are higher than those of Y2DyN@C80 and YDy2N@C80, indicating that Y2Dy@C80 and YDy2@C80 exhibit greater stability. It is important to note that the arc-discharge synthesis method involves conditions that are far from equilibrium. These conditions, referred to as non-equilibrium plasma by Osawa [51], prioritize structural and flux parameters over energy parameters. In previous studies, auxiliary parameters, such as activation energies and stochastic descriptors, have been utilized to predict the formation of fullerenes [52]. In the case of Y3@C80, Y2Dy@C80, and YDy2@C80, the analysis based on stochastic descriptors may offer valuable insights. This aspect will be explored in our future research.
Y 2 Dy + C 80 Y 2 Dy @ C 80 Δ E b ( 1 ) = 281 kcal / mol
YDy 2 + C 80 YDy 2 @ C 80 Δ E b ( 2 ) = 273 kcal / mol
Y 2 DyN + C 80 Y 2 DyN @ C 80 Δ E b ( 3 ) = 269 kcal / mol
YDy 2 N + C 80 YDy 2 N @ C 80 Δ E b ( 4 ) = 257 kcal / mol .
Chemical oxidation plays a crucial role in the extraction of Y2Dy@C80 and YDy2@C80. Attempting direct extraction using conventional fullerene solvents such as toluene and CS2 proved unsuccessful for these mixed rare-earth tri-metallofullerenes. This scenario echoes our earlier findings with Tm3@C80 [43]. The inference drawn is that Y2Dy@C80 and YDy2@C80 tend to undergo oxidation, and their resulting cations demonstrate chemical stability. Subsequently, we computed the ionization energy of Y2Dy@C80 and YDy2@C80, and observed that they exhibit relatively low ionization energy values. Specifically, the ionization energy of Y2Dy@C80 is 5.49 eV, while that of YDy2@C80 is 5.47 eV. These values are significantly smaller than the calculated ionization energies of C60 (7.76 eV) and C70 (7.56 eV). It has been reported that Li@C60 can undergo oxidation to form L i @ C 60 + , leading to the creation of various stable ionic compounds [53,54,55]. Notably, the calculated ionization energy of Y2Dy@C80 and YDy2@C80 is comparable to that of Li@C60 (5.73 eV), computed at the same level. Consequently, it can be inferred that Y2Dy@C80 and YDy2@C80 molecules may undergo chemical oxidation to produce their respective cations, namely Y 2 D y @ C 80 + and Y D y 2 @ C 80 + . Figure 2 presents the DFT-optimized structures of Y 2 D y @ C 80 + and Y D y 2 @ C 80 + . For comparison, the DFT-optimized structures of Y2DyN@C80 and YDy2N@C80 are also depicted in Figure 2. The comparison reveals that the structures of Y 2 D y @ C 80 + and Y D y 2 @ C 80 + closely resemble those of Y2DyN@C80 and YDy2N@C80, respectively. The metal–metal distances are around 3.5 angstrom in these cations and neutral molecules.
The low ionization energy exhibited by Y2Dy@C80 and YDy2@C80 can be elucidated by examining the energy of molecular orbitals. We carried out calculations on the clusters of Y2Dy, YDy2, and the fullerene C80. Figure 3a displays the calculated molecular orbital diagrams. The LUMO/LUMO+1/LUMO+2 orbitals of the C80 molecule are nearly degenerate, while there is a significantly large energy gap between LUMO + 2 and LUMO + 3. As a result, the C80 cage has a preference to accept six electrons in order to achieve a stable electronic configuration. In the case of the Y2Dy, YDy2 clusters, there are seven electrons with relatively high energy levels in the frontier molecular. Upon encapsulation of the metal cluster into the C80 cage, six electrons are transferred to the LUMO/LUMO + 1/LUMO + 2 orbitals of C80, leaving one unpaired electron on the metal cluster. In the case of the bare Y2Dy and YDy2 clusters, there exist molecular orbitals that are associated with three-center two-electron bonding (encircled with dotted lines in Figure 3a). These specific molecular orbitals retain their bonding characteristics when the metal cluster is enclosed within C80 due to their comparably low energy levels.
The bond lengths for Dy-Y(1), Dy-Y(2), and Y(1)-Y(2) in bare Y2Dy cluster are 3.21, 3.21, and 3.18 angstrom, respectively, which are much shorter than those in the metallofullerene Y2Dy@C80. We carried out a Wisberg bond order analysis for Y2Dy, YDy2, Y2Dy@C80 and YDy2@C80. In Y2Dy, the bond orders for Dy-Y(1), Dy-Y(2), and Y(1)-Y(2) are 1.76, 1.76, and 2.05, respectively, indicating a strong bonding interaction between the metal atoms. However, in Y2Dy@C80, the bond orders for Dy-Y(1), Dy-Y(2), and Y(1)-Y(2) decrease to 0.56, 0.62, and 0.75, respectively. This weakening of the metal atom bonding is a result of electron transfer from the metal cluster to the fullerene cage. At the same time, the repulsion between the metal ions becomes stronger in the metallofullerene compared to the bare metal cluster. A similar trend is observed for YDy2 and YDy2@C80. In YDy2, the bond orders for Dy(1)-Dy(2), Dy(1)-Y, and Dy(2)-Y are 1.64, 1.94, and 1.94, respectively. These bond orders decrease to 0.54, 0.69, and 0.67 in YDy2@C80.
Figure 3b,c portray the calculated molecular orbitals for Y2Dy@C80 and YDy2@C80. In both instances, the HOMO for the alpha spin possesses a relatively high energy level. Figure 3d displays the calculated spin density distribution for Y2Dy@C80 and YDy2@C80. They exhibit a spatial distribution that is comparable to the alpha HOMO, signifying that the alpha HOMO primarily corresponds to the unpaired electron in the molecule. The high energy level of the alpha HOMO makes the molecule prone to oxidation into a cation. Upon careful analysis of the alpha HOMO’s shape, it has been determined that it is a bonding orbital. The removal of an electron from this orbital would result in a reduction of the bonding strength between the metal atoms. Consequently, the cations Y 2 D y @ C 80 + and Y D y 2 @ C 80 + exhibit longer metal–metal distances compared to the neutral molecules Y2Dy@C80 and YDy2@C80 (Figure 2).
As previously mentioned, the molecular structures of Y 2 D y @ C 80 + and Y D y 2 @ C 80 + closely mirror those of Y2DyN@C80 and YDy2N@C80, respectively. The metal–metal distances in Y 2 D y @ C 80 + and Y D y 2 @ C 80 + closely resemble those in Y2DyN@C80 and YDy2N@C80. Our calculations further reveal that they share similar electronic structures. Figure 4a,b illustrate the molecular orbital energy diagrams for Y 2 D y @ C 80 + and Y2DyN@C80. Notably, the HOMO and LUMO of Y 2 D y @ C 80 + exhibit strikingly similar shapes to that of Y2DyN@C80. Meanwhile, the calculated HOMO-LUMO gap of Y 2 D y @ C 80 + (2.80 eV) is identical to that of Y2DyN@C80 (2.80 eV). Among all the reported metallofullerenes, M3N@C80 represents a type with a notably large HOMO-LUMO gap. DFT calculations in this study unveil that Y 2 D y @ C 80 + also possesses a substantial HOMO-LUMO gap, suggesting high stability for this cation. It is noteworthy that the neutral Y2Dy@C80 exhibits a relatively small HOMO-LUMO gap (1.47 eV using the PBE0 functional, and 0.30 eV using the PBE functional), indicating its chemical reactivity. The significant increase in the HOMO-LUMO gap for Y 2 D y @ C 80 + enhances the chemical stability of Y2Dy@C80 upon oxidation.
It is noteworthy that the ELF distribution of Y 2 D y @ C 80 + closely resembles that of Y2DyN@C80 (Figure 4c). In the case of Y 2 D y @ C 80 + , the ELF distribution at the molecular center indicates the presence of covalent bonding among the three metal atoms, specifically forming a three-center metal–metal bond. This bond corresponds to the HOMO-3 molecular orbital, as depicted in Figure 4a. For Y2DyN@C80, where the formal charge of the metal atom is +3, strong electrostatic repulsion exists between the metal ions. The central N3− ion acts as a mediator, mitigating the repulsion between the metal ions. In the bare Y2Dy cluster, the bond lengths for Dy-Y(1), Dy-Y(2), and Y(1)-Y(2) are 3.21, 3.21, and 3.18 angstrom, respectively. However, in Y 2 D y @ C 80 + , the bond lengths for Dy-Y(1), Dy-Y(2), and Y(1)-Y(2) increase to 3.51, 3.51, and 3.51 angstrom, respectively, indicating strong repulsion between these metal ions. A similar trend is observed for YDy2 and YDy2@C80. In YDy2, the bond lengths for Dy(1)-Dy(2), Dy(1)-Y, and Dy(2)-Y are 3.24, 3.21, and 3.21 angstrom, respectively. These bond lengths increase to 3.51, 3.49, and 3.50 angstrom in Y D y 2 @ C 80 + . In the case of Y 2 D y @ C 80 + , the three-center σ bond serves a role akin to that of the N3− ion in Y2DyN@C80. The three-center σ bond can mitigate the repulsion between the metal ions and allows three metal atoms to be accommodated within the C80 cage without the need for a nonmetal mediator. Thus, the three-center σ bond emerges as a crucial factor contributing to the stabilization of the tri-metallofullerene cation. Furthermore, our calculations indicate that the electronic structure of Y D y 2 @ C 80 + closely mirrors that of YDy2N@C80. They exhibit analogous frontier molecular orbitals and comparable ELF distribution at the molecular center (Figure 5). Consequently, the stability of Y D y 2 @ C 80 + can be elucidated in a manner similar to that of Y 2 D y @ C 80 + .
The magnetic characteristics of rare-earth metallofullerenes are intriguing. Some of them are formed in high spin states [56,57]. There have been reports highlighting the single-molecule magnet behavior of the clusterfullerenes Y2DyN@C80 and YDy2N@C80 [58]. Given that the cations Y 2 D y @ C 80 + and Y D y 2 @ C 80 + share similar geometrical and electronic structures with Y2DyN@C80 and YDy2N@C80, respectively, these cations emerge as promising candidates for single-molecule magnets. Exploring the separation of these cations in future studies would constitute valuable and important research.

3. Experimental and Theoretical Methods

The metallofullerenes were synthesized using the arc-discharge technique. The anode was made of a carbon rod, while the cathode was a hollow carbon rod filled with a mixture of Y2O3, Dy2O3 (molar ratio of Y:Dy = 2:1), and graphite powders. Prior to the discharge, the cathode was heated by connecting it to the anode and applying a current of 120 A for an hour under a dynamic vacuum. The actual arc-discharge process occurred at a current of 120 A in a 200 Torr helium environment, with a distance of approximately 1 cm between the anode and cathode. The raw soot generated from the arc was subjected to oxidation using AgSbF6 in dichloromethane within a nitrogen-filled glove box for a duration of 24 h. Following the oxidation process, the solution was separated from the insoluble soot residue through centrifugation and filtration. No chromatographic purification was carried out. The sample was then analyzed using a mass spectrometer (AB SCIEX 5800 MALDI TOF/TOF, Toronto, Canada) to obtain a mass spectrum, with the analysis conducted in the positive mode without using any matrix.
We carried out DFT calculations to examine the structure and properties of the metallofullerenes. To optimize the structure and determine the molecule’s energy, the PBE0 functional [59] was employed. The carbon and nitrogen atoms were treated using the 6-31G(d) basis set [60]. The Y and Dy atoms were considered using the SDD pseudopotentials and the corresponding basis sets. The 6-311G(d) basis set was used for single-point energy calculations [61]. To investigate the bonding properties of the molecules, the MULTIWFN program [62] was utilized to perform electron localization function (ELF) analysis [63] and Wiberg bond order analysis on a Lowdin orthogonalized basis. All DFT calculations were conducted using Gaussian16 version A03 [64]. The visualization of the calculation results was accomplished using the Visual Molecular Dynamics (VMD, version 1.9.3) software [65].

4. Conclusions

We have conducted a comprehensive investigation into the stability and electronic properties of mixed rare-earth tri-metallofullerenes, specifically YxDy3−x@C80 (x = 1 or 2), through a combined experimental and theoretical study. Our findings indicate that the chemical oxidation of the arc-discharge produced soot can result in the formation of cations, namely Y 2 D y @ C 80 + and Y D y 2 @ C 80 + . Through DFT calculations, we have determined that the tri-metallofullerenes YxDy3−x@C80 (x = 1 or 2) possess a low oxidation potential, significantly lower than that of the other fullerenes like C60 and C70. This low oxidation potential is attributed to the relatively high energy level of the singly occupied orbital alpha HOMO. Furthermore, the oxidized species exhibit a large HOMO-LUMO gap similar to that of YxDy3−xN@C80, highlighting their high chemical stability. Theoretical studies have revealed the presence of a three-center two-electron metal–metal bond at the center of Y 2 D y @ C 80 + and Y D y 2 @ C 80 + . This unique multi-center bond helps alleviate the electrostatic repulsion between the metal ions, thus contributing to the overall stability of the cations.

Author Contributions

Conceptualization Z.W.; methodology, Y.W. and Z.Z.; software, Y.W. and Z.Z.; validation, Y.W. and Z.Z.; formal analysis, Y.W. and Z.Z.; investigation, Y.W. and Z.Z.; resources, Z.W.; data curation, Z.Z. and Z.W.; writing—original draft preparation, Z.W.; writing—review and editing, Z.W.; visualization, Y.W.; supervision, Z.W.; project administration, Z.W.; funding acquisition, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Beijing Municipal Natural Science Foundation (grant no. 2212030) and the National Natural Science Foundation of China (grant no. 22175199).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data reported in this study are available upon request by contact with the corresponding author. The data are not publicly available due to technical limitations (with a total size of a few GB).

Acknowledgments

We are grateful for the support of the Public Computing Cloud at Renmin University of China.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shinohara, H. Endohedral metallofullerenes. Rep. Prog. Phys. 2000, 63, 843. [Google Scholar] [CrossRef]
  2. Chaur, M.N.; Melin, F.; Ortiz, A.L.; Echegoyen, L. Chemical, electrochemical, and structural properties of endohedral metallofullerenes. Angew. Chem. Int. Ed. 2009, 48, 7514–7538. [Google Scholar] [CrossRef]
  3. Cong, H.; Yu, B.; Akasaka, T.; Lu, X. Endohedral metallofullerenes: An unconventional core–shell coordination union. Coord. Chem. Rev. 2013, 257, 2880–2898. [Google Scholar]
  4. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  5. Wang, T.; Wang, C. Endohedral Metallofullerenes Based on Spherical Ih-C80 Cage: Molecular Structures and Paramagnetic Properties. Acc. Chem. Res. 2014, 47, 450–458. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, S.; Wei, T.; Jin, F. When metal clusters meet carbon cages: Endohedral clusterfullerenes. Chem. Soc. Rev. 2017, 46, 5005–5058. [Google Scholar] [CrossRef] [PubMed]
  7. Dang, J.S.; Wang, W.W.; Zheng, J.J.; Zhao, X.; Nagase, S. Fused-Pentagon-Configuration-Dependent Electron Transfer of Monotitanium-Encapsulated Fullerenes. Inorg. Chem. 2017, 56, 6890–6896. [Google Scholar] [CrossRef]
  8. 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]
  9. Shen, W.; Hu, S.; Lu, X. Endohedral metallofullerenes: New structures and unseen phenomena. Chem. Eur. J. 2020, 26, 5748–5757. [Google Scholar] [CrossRef]
  10. Li, M.; Zhao, R.; Dang, J.; Zhao, X. Theoretical study on the stabilities, electronic structures, and reaction and formation mechanisms of fullerenes and endohedral metallofullerenes. Coord. Chem. Rev. 2022, 471, 214762. [Google Scholar] [CrossRef]
  11. Shen, W.; Bao, L.; Lu, X. Endohedral Metallofullerenes: An Ideal Platform of Sub-Nano Chemistry. Chin. J. Chem. 2022, 40, 275–284. [Google Scholar] [CrossRef]
  12. Hu, Z.; Wang, Y.; Ullah, A.; Gutiérrez-Finol, G.M.; Bedoya-Pinto, A.; Gargiani, P.; Shi, D.; Yang, S.; Shi, Z.; Gaita-Ariño, A.; et al. High-temperature magnetic blocking in a monometallic dysprosium azafullerene single-molecule magnet. Chem 2023, 9, 3613–3622. [Google Scholar] [CrossRef]
  13. Zhang, X.; Wang, Y.; Morales-Martínez, R.; Zhong, J.; de Graaf, C.; Rodríguez-Fortea, A.; Poblet, J.M.; Echegoyen, L.; Feng, L.; Chen, N. U2@Ih(7)-C80: Crystallographic Characterization of a Long-Sought Dimetallic Actinide Endohedral Fullerene. J. Am. Chem. Soc. 2018, 140, 3907–3915. [Google Scholar] [CrossRef] [PubMed]
  14. Li, X.; Roselló, Y.; Yao, Y.R.; Zhuang, J.; Zhang, X.; Rodríguez-Fortea, A.; de Graaf, C.; Echegoyen, L.; Poblet, J.M.; Chen, N. U2N@Ih(7)-C80: Fullerene Cage Encapsulating an Unsymmetrical U (IV) = N = U (V) Cluster. Chem. Sci. 2021, 12, 282–292. [Google Scholar] [CrossRef]
  15. Junghans, K.; Rosenkranz, M.; Popov, A.A. Sc3CH@C80: Selective 13C enrichment of the central carbon atom. Chem. Commun. 2016, 52, 6561–6564. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, C.; Tan, K.; Nie, M.; Lu, Y.; Zhang, J.; Wang, C.; Lu, X.; Wang, T. Scandium Tetrahedron Supported by H Anion and CN Pentaanion inside Fullerene C80. Inorg. Chem. 2020, 59, 8284–8290. [Google Scholar] [CrossRef]
  17. Fuertes-Espinosa, C.; Gómez-Torres, A.; Morales-Martínez, R.; Rodríguez-Fortea, A.; García-Simón, C.; Gándara, F.; Imaz, I.; Juanhuix, J.; Maspoch, D.; Poblet, J.M.; et al. Purification of Uranium-based Endohedral Metallofullerenes (EMFs) by Selective Supramolecular Encapsulation and Release. Angew. Chem. 2018, 130, 11464–11469. [Google Scholar] [CrossRef]
  18. Wei, Z.; Yang, L.; Ji, J.; Hou, Q.; Li, L.; Jin, P. Undiscovered Effect of C ↔ N Interchange Inside the Metal Carbonitride Clusterfullerenes: A Density Functional Theory Investigation. Inorg. Chem. 2020, 59, 6518–6527. [Google Scholar] [CrossRef]
  19. Wei, T.; Wang, S.; Lu, X.; Tan, Y.; Huang, J.; Liu, F.; Li, Q.; Xie, S.; Yang, S. Entrapping a group-VB transition metal, vanadium, within an endohedral metallofullerene: V x Sc3–x N@ I h-C80 (x= 1, 2). J. Am. Chem. Soc. 2016, 138, 207–214. [Google Scholar] [CrossRef]
  20. Takata, M.; Umeda, B.; Nishibori, E.; Sakata, M.; Saitot, Y.; Ohno, M.; Shinohara, H. Confirmation by X-ray diffraction of the endohedral nature of the metallofullerene Y@C82. Nature 1995, 377, 46–49. [Google Scholar] [CrossRef]
  21. Shinohara, H.; Sato, H.; Saito, Y.; Ohkohchi, M.; Ando, Y. Mass spectroscopic and ESR characterization of soluble yttrium-containing metallofullerenes YC82 and Y2C82. J. Phys. Chem. 2002, 96, 3571–3573. [Google Scholar] [CrossRef]
  22. Xu, D.; Jiang, Y.; Wang, Y.; Zhou, T.; Shi, Z.; Omachi, H.; Shinohara, H.; Sun, B.; Wang, Z. Turning on the Near-Infrared Photoluminescence of Erbium Metallofullerenes by Covalent Modification. Inorg. Chem. 2019, 58, 14325–14330. [Google Scholar] [CrossRef]
  23. Lu, X.; Akasaka, T.; Nagase, S. Chemistry of endohedral metallofullerenes: The role of metals. Chem. Commun. 2011, 47, 5942–5957. [Google Scholar] [CrossRef] [PubMed]
  24. Yamada, M.; Liu, M.T.; Nagase, S.; Akasaka, T. New horizons in chemical functionalization of endohedral metallofullerenes. Molecules 2020, 25, 3626. [Google Scholar] [CrossRef]
  25. Akasaka, T.; Wakahara, T.; Nagase, S.; Kobayashi, K.; Waelchli, M.; Yamamoto, K.; Kondo, M.; Shirakura, S.; Okubo, S.; Maeda, Y.; et al. La@ C82 anion. An unusually stable metallofullerene. J. Am. Chem. Soc. 2000, 122, 9316–9317. [Google Scholar] [CrossRef]
  26. Feng, L.; Nakahodo, T.; Wakahara, T.; Tsuchiya, T.; Maeda, Y.; Akasaka, T.; Kato, T.; Horn, E.; Yoza, K.; Mizorogi, N.; et al. A singly bonded derivative of endohedral metallofullerene: La@ C82CBr (COOC2H5) 2. J. Am. Chem. Soc. 2005, 127, 17136–17137. [Google Scholar] [CrossRef]
  27. García-Hernández, D.A.; Manchado, A.; Cataldo, F. Hydrogenation of [Li@C60]PF6: A comparison with fulleranes derived from C60. Fuller. Nanotub. Carbon Nanostruct. 2022, 30, 1245–1251. [Google Scholar]
  28. Kobayashi, K.; Nagase, S.; Akasaka, T. A theoretical study of C80 and La2@C80. Chem. Phys. Lett. 1995, 245, 230–236. [Google Scholar] [CrossRef]
  29. Wang, Z.; Kitaura, R.; Shinohara, H. Metal-dependent stability of pristine and functionalized unconventional dimetallofullerene M2@Ih-C80. J. Phys. Chem. C 2014, 118, 13953–13958. [Google Scholar] [CrossRef]
  30. Jiang, Y.; Li, Z.; Zhang, J.; Wang, Z. Endohedral metallofullerene M2@C80: A new class of magnetic superhalogen. J. Phys. Chem. C 2019, 124, 2131–2136. [Google Scholar] [CrossRef]
  31. Liu, F.; Velkos, G.; Krylov, D.S.; Spree, L.; Zalibera, M.; Ray, R.; Samoylova, N.A.; Chen, C.H.; Rosenkranz, M.; Schiemenz, S.; et al. Air-stable redox-active nanomagnets with lanthanide spins radical-bridged by a metal–metal bond. Nat. Commun. 2019, 10, 571. [Google Scholar] [CrossRef] [PubMed]
  32. Krause, M.; Dunsch, L. Isolation and characterisation of two Sc3N@C80 isomers. ChemPhysChem 2004, 5, 1445–1449. [Google Scholar] [CrossRef]
  33. Kurihara, H.; Iiduka, Y.; Rubin, Y.; Waelchli, M.; Mizorogi, N.; Slanina, Z.; Tsuchiya, T.; Nagase, S.; Akasaka, T. Unexpected formation of a Sc3C2@C80 bisfulleroid derivative. J. Am. Chem. Soc. 2012, 134, 4092–4095. [Google Scholar] [CrossRef]
  34. Wang, T.S.; Chen, N.; Xiang, J.F.; Li, B.; Wu, J.Y.; Xu, W.; Jiang, L.; Tan, K.; Shu, C.Y.; Lu, X.; et al. Russian-doll-type metal carbide endofullerene: Synthesis, isolation, and characterization of Sc4C2@C80. J. Am. Chem. Soc. 2009, 131, 16646–16647. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, P.; Shen, W.; Bao, L.; Pan, C.; Slanina, Z.; Lu, X. Trapping an unprecedented Ti3C3 unit inside the icosahedral C80 fullerene: A crystallographic survey. Chem. Sci. 2019, 10, 10925–10930. [Google Scholar] [CrossRef]
  36. Jin, F.; Xin, J.; Guan, R.; Xie, X.M.; Chen, M.; Zhang, Q.; Popov, A.A.; Xie, S.Y.; Yang, S. Stabilizing a three-center single-electron metal–metal bond in a fullerene cage. Chem. Sci. 2021, 12, 6890–6895. [Google Scholar] [CrossRef] [PubMed]
  37. Tagmatarchis, N.; Aslanis, E.; Prassides, K.; Shinohara, H. Mono-, di- and trierbium endohedral metallofullerenes: Production, separation, isolation, and spectroscopic study. Chem. Mater. 2001, 13, 2374–2379. [Google Scholar] [CrossRef]
  38. Yang, S.; Dunsch, L. Di-and tridysprosium endohedral metallofullerenes with cages from C94 to C100. Angew. Chem. Int. Ed. 2006, 45, 1299–1302. [Google Scholar] [CrossRef]
  39. Guo, Y.J.; Zheng, H.; Yang, T.; Nagase, S.; Zhao, X. Theoretical Insight into the Ambiguous Endohedral Metallofullerene Er3C74: Covalent Interactions among Three Lanthanide Atoms. Inorg. Chem. 2015, 54, 8066–8076. [Google Scholar] [CrossRef]
  40. Lian, Y.; Shi, Z.; Zhou, X.; Gu, Z. Different extraction Behaviors between divalent and trivalent endohedral metallofullerenes. Chem. Mater. 2004, 16, 1704–1714. [Google Scholar] [CrossRef]
  41. Popov, A.A.; Zhang, L.; Dunsch, L. A pseudoatom in a cage: Trimetallofullerene Y3@C80 mimics Y3N@C80 with nitrogen substituted by a pseudoatom. ACS Nano 2010, 4, 795–802. [Google Scholar] [CrossRef]
  42. Xu, W.; Feng, L.; Calvaresi, M.; Liu, J.; Liu, Y.; Niu, B.; Shi, Z.; Lian, Y.; Zerbetto, F. An Experimentally Observed Trimetallofullerene Sm3@Ih-C80: Encapsulation of Three Metal Atoms in a Cage without a Nonmetallic Mediator. J. Am. Chem. Soc. 2013, 135, 4187–4190. [Google Scholar] [CrossRef]
  43. Jiang, Y.; Li, Z.; Wu, Y.; Wang, Z. Ln3@C80+(Ln= lanthanide): A new class of stable metallofullerene cations with multicenter metal–metal bonding in the sub-nanometer confined space. Inorg. Chem. Front. 2022, 9, 2173–2181. [Google Scholar] [CrossRef]
  44. Wang, Y.; Ma, L.; Mu, L.; Ren, J.; Kong, X. A systematic study on the generation of multimetallic lanthanide fullerene ions by laser ablation mass spectrometry. Rapid Commun. Mass Spectrom. 2018, 32, 1396–1402. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, Y.; Ren, J.; Mu, L.; Kong, X. Metallofullerene ions of Lun C 2 m + (1⩽n⩽9, 50⩽2m⩽198) generated by laser ablation of graphene/LuCl3. Int. J. Mass Spectrom. 2017, 422, 105–110. [Google Scholar] [CrossRef]
  46. Mu, L.; Yang, S.; Feng, R.; Kong, X. Encapsulation of Platinum in Fullerenes: Is That Possible? Inorg. Chem. 2017, 56, 6035–6038. [Google Scholar] [CrossRef]
  47. Kong, X.; Bao, X. Formation of endohedral metallofullerene (EMF) ions of (M = La, Y, n ⩽ 6, 50 ⩽ 2 m ⩽ 194) in the laser ablation process with graphene as precursor. Rapid Commun. Mass Spectrom. 2017, 31, 865–872. [Google Scholar] [CrossRef]
  48. Sabirov, D.S.; Tukhbatullina, A.A.; Shepelevich, I.S. Digitalizing Structure–Symmetry Relations at the Formation of Endofullerenes in Terms of Information Entropy Formalism. Symmetry 2022, 14, 1800. [Google Scholar] [CrossRef]
  49. Popov, A.A.; Dunsch, L. Structure, Stability, and Cluster-Cage Interactions in Nitride Clusterfullerenes M3N@C2n(M = Sc, Y; 2n = 68-98): A Density Functional Theory Study. J. Am. Chem. Soc. 2007, 129, 11835–11849. [Google Scholar] [CrossRef]
  50. Jin, P.; Tang, C.; Chen, Z. Carbon atoms trapped in cages: Metal carbide clusterfullerenes. Coord. Chem. Rev. 2014, 270, 89–111. [Google Scholar] [CrossRef]
  51. Osawa, E. Formation Mechanism of C60 under Nonequilibrium and Irreversible Conditions—An Annotation. Fuller. Nanotub. Carbon Nanostruct. 2012, 20, 299–309. [Google Scholar] [CrossRef]
  52. Sabirov, D.S.; Terentyev, A.O.; Sokolov, V.I. Activation energies and information entropies of helium penetration through fullerene walls. Insights into the formation of endofullerenes nX@C60/70 (n = 1 and 2) from the information entropy approach. RSC Adv. 2016, 6, 72230–72237. [Google Scholar]
  53. Aoyagi, S.; Nishibori, E.; Sawa, H.; Sugimoto, K.; Takata, M.; Miyata, Y.; Kitaura, R.; Shinohara, H.; Okada, H.; Sakai, T.; et al. A layered ionic crystal of polar Li@C60 superatoms. Nat. Chem. 2010, 2, 678–683. [Google Scholar] [CrossRef]
  54. Aoyagi, S.; Sado, Y.; Nishibori, E.; Sawa, H.; Okada, H.; Tobita, H.; Kasama, Y.; Kitaura, R.; Shinohara, H. Rock-Salt-Type Crystal of Thermally Contracted C60 with Encapsulated Lithium Cation. Angew. Chem. Int. Ed. 2012, 124, 3433–3437. [Google Scholar] [CrossRef]
  55. Okada, H.; Komuro, T.; Sakai, T.; Matsuo, Y.; Ono, Y.; Omote, K.; Yokoo, K.; Kawachi, K.; Kasama, Y.; Ono, S.; et al. Preparation of endohedral fullerene containing lithium (Li@C60) and isolation as pure hexafluorophosphate salt ([Li+@C60][ P F 6 ]). RSC Adv. 2012, 2, 10624–10631. [Google Scholar]
  56. Kandrashkin, Y.E.; Zaripov, R.B. Low-temperature motion of the scandium bimetal in endofullerene Sc@C80(CH2Ph). Phys. Chem. Chem. Phys. 2023, 25, 31493–31499. [Google Scholar] [CrossRef]
  57. Bhusal, S.; Baruah, T.; Yamamoto, Y.; Zope, R.R. Electronic structure calculation of vanadium-and scandium-based endohedral fullerenes VSc2N@C2n(2n = 70, 76, 78, 80). Int. J. Quantum. Chem. 2018, 118, e25785. [Google Scholar] [CrossRef]
  58. Hao, Y.; Velkos, G.; Schiemenz, S.; Rosenkranz, M.; Wang, Y.; Büchner, B.; Avdoshenko, S.M.; Popov, A.A.; Liu, F. Using internal strain and mass to modulate Dy⋯Dy coupling and relaxation of magnetization in heterobimetallic metallofullerenes DyM2N@C80 and Dy2MN@C80 (M = Sc, Y, La, Lu). Inorg. Chem. Front. 2023, 10, 468–484. [Google Scholar] [CrossRef]
  59. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  60. Hariharan, P.C.; Pople, J.A. The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  61. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  62. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  63. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. ELF: The electron localization function. Angew. Chem. Int. Ed. Engl. 1997, 36, 1808–1832. [Google Scholar] [CrossRef]
  64. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Revision A.03; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  65. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Mass spectrum for the oxidized metallofullerenes. The signals from Y3@C80, Y2Dy@C80 and YDy2@C80 are observed. (b) The experimental and calculated isotope distributions of the samples.
Figure 1. (a) Mass spectrum for the oxidized metallofullerenes. The signals from Y3@C80, Y2Dy@C80 and YDy2@C80 are observed. (b) The experimental and calculated isotope distributions of the samples.
Molecules 29 00447 g001
Figure 2. The optimized structures of (a) Y2Dy@C80, Y 2 D y @ C 80 + , Y2DyN@C80 and (b) YDy2@C80, Y D y 2 @ C 80 + , YDy2N@C80. The metal–metal distances are also shown. The C, N, Y, and Dy atoms are denoted by gray, blue, green, and brown colors, respectively.
Figure 2. The optimized structures of (a) Y2Dy@C80, Y 2 D y @ C 80 + , Y2DyN@C80 and (b) YDy2@C80, Y D y 2 @ C 80 + , YDy2N@C80. The metal–metal distances are also shown. The C, N, Y, and Dy atoms are denoted by gray, blue, green, and brown colors, respectively.
Molecules 29 00447 g002
Figure 3. (a) The calculated molecular orbital energy level of C80 compared to that of Y2Dy and YDy2 clusters. The molecular orbitals corresponding to three-center two-electron metal–metal bonding in the clusters are encircled with dotted lines. The spatial distribution of the three-center two-electron metal–metal bonding orbital for the Y2Dy and YDy2 clusters is shown. (b,c) The molecular orbital energy diagrams for Y2Dy@C80 and YDy2@C80. (d) The calculated spin density distribution for Y2Dy2@C80 and YDy2@C80 (isovalue = 0.0004).
Figure 3. (a) The calculated molecular orbital energy level of C80 compared to that of Y2Dy and YDy2 clusters. The molecular orbitals corresponding to three-center two-electron metal–metal bonding in the clusters are encircled with dotted lines. The spatial distribution of the three-center two-electron metal–metal bonding orbital for the Y2Dy and YDy2 clusters is shown. (b,c) The molecular orbital energy diagrams for Y2Dy@C80 and YDy2@C80. (d) The calculated spin density distribution for Y2Dy2@C80 and YDy2@C80 (isovalue = 0.0004).
Molecules 29 00447 g003
Figure 4. The molecular orbital energy diagrams for (a) Y2Dy@ C 80 + and (b) Y2DyN@C80. HOMO-3 in (a) corresponds to a three-center σ bond. (c) Three-dimensional ELF isosurface (left, isovalue = 0.6) and color-filled two-dimensional ELF maps at the metal cluster plane (right) for Y2Dy@ C 80 + and Y2DyN@C80.
Figure 4. The molecular orbital energy diagrams for (a) Y2Dy@ C 80 + and (b) Y2DyN@C80. HOMO-3 in (a) corresponds to a three-center σ bond. (c) Three-dimensional ELF isosurface (left, isovalue = 0.6) and color-filled two-dimensional ELF maps at the metal cluster plane (right) for Y2Dy@ C 80 + and Y2DyN@C80.
Molecules 29 00447 g004
Figure 5. The molecular orbital energy diagrams for (a) Y D y 2 @ C 80 + and (b) YDy2N@C80. HOMO-3 in (a) corresponds to a three-center σ bond. (c) Three-dimensional ELF isosurface (left, isovalue = 0.6) and color-filled two-dimensional ELF maps at the metal cluster plane (right) for YDy2@ C 80 + and YDy2N@C80.
Figure 5. The molecular orbital energy diagrams for (a) Y D y 2 @ C 80 + and (b) YDy2N@C80. HOMO-3 in (a) corresponds to a three-center σ bond. (c) Three-dimensional ELF isosurface (left, isovalue = 0.6) and color-filled two-dimensional ELF maps at the metal cluster plane (right) for YDy2@ C 80 + and YDy2N@C80.
Molecules 29 00447 g005
Table 1. The calculated relative energy (kcal/mol) of Y2Dy@C80 with different fullerene cages. The spiral number of the fullerene cage is included.
Table 1. The calculated relative energy (kcal/mol) of Y2Dy@C80 with different fullerene cages. The spiral number of the fullerene cage is included.
Molecular FormulaRelative Energy
Y2Dy@Ih-31924C800.00
Y2Dy@D5h-31923C8012.04
Y2Dy@C2v-31922C8017.13
Y2Dy@C1-28325C8019.50
Y2Dy@C2-29591C8024.54
Y2Dy@C2v-31920C8025.58
Y2Dy@C1-31876C8028.06
Y2Dy@C2-28319C8029.20
Y2Dy@C1-28324C8030.50
Table 2. The calculated relative energy (kcal/mol) of YDy2@C80 with different fullerene cages.
Table 2. The calculated relative energy (kcal/mol) of YDy2@C80 with different fullerene cages.
Molecular FormulaRelative Energy
YDy2@Ih-31924C800.00
YDy2@D5h-31923C8012.08
YDy2@C2v-31922C8018.67
YDy2@C1-28325C8023.05
YDy2@C2-29591C8027.81
YDy2@C2v-31920C8028.11
YDy2@C1-31876C8029.76
YDy2@C2-28319C8030.49
YDy2@C1-28314C8032.32
Table 3. The calculated relative energy (kcal/mol) of Y2Dy@Ih-C80 and Y2DyC2@C78 with different fullerene cages. The spiral number of the fullerene cage is included.
Table 3. The calculated relative energy (kcal/mol) of Y2Dy@Ih-C80 and Y2DyC2@C78 with different fullerene cages. The spiral number of the fullerene cage is included.
Molecular FormulaRelative Energy
Y2Dy@Ih-C800.00
Y2DyC2@C2-22010C7842.95
Y2DyC2@C1-21975C7859.17
Y2DyC2@C2v-24107C7859.25
Y2DyC2@C2v-24088C7871.57
Y2DyC2@D3h-24109C7890.81
Table 4. The calculated relative energy (kcal/mol) of YDy2@Ih-C80 and YDy2C2@C78 with different fullerene cages. The spiral number of the fullerene cage is included.
Table 4. The calculated relative energy (kcal/mol) of YDy2@Ih-C80 and YDy2C2@C78 with different fullerene cages. The spiral number of the fullerene cage is included.
Molecular FormulaRelative Energy
YDy2@Ih-C800.00
YDy2C2@C2-22010C7849.10
YDy2C2@C2v-24107C7863.04
YDy2C2@C1-21975C7865.24
YDy2C2@C2v-24088C7878.04
YDy2C2@D3h-24109C7895.57
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

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. https://doi.org/10.3390/molecules29020447

AMA Style

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(2):447. https://doi.org/10.3390/molecules29020447

Chicago/Turabian Style

Wu, Yabei, Zhonghao Zhou, and Zhiyong Wang. 2024. "Stability and Electronic Properties of Mixed Rare-Earth Tri-Metallofullerenes YxDy3-x@C80 (x = 1 or 2)" Molecules 29, no. 2: 447. https://doi.org/10.3390/molecules29020447

Article Metrics

Back to TopTop