Abstract
The empty space inside a fullerene cage can be filled with a variety of species, including metal dimers. Encapsulation of Sc2, Y2, or lanthanide dimers leads to dimetallofullerenes featuring metal–metal bonding molecular orbital. Such an orbital can be either HOMO or LUMO of the dimetallofullerene molecule. In certain cases, single-occupied metal–metal bonding orbital can be also stabilized. This review is focused on redox processes involving variation of the electron population of metal–metal bonding orbitals in dimetallofullerenes.
Introduction
The encapsulation of metal atoms by carbon cages in endohedral metallofullerenes (EMFs) leads to a plethora of interesting chemical and physical phenomena [1–5]. High chemical and thermal stability of fullerene cage protects endohedral entities from the environment and can stabilize unusual species, which cannot exist otherwise. Metal atoms enclosed inside a fullerene transfer their valence electrons to the carbon cage, resulting in “salts” with cationic metals and anionic fullerene cages. Electrochemistry has been traditionally used as a relatively simple and yet very powerful technique to study electronic structures of EMFs [1,6•].
Empty fullerenes are good electron acceptors and undergo multiple single-electron redox steps in solutions [7]. Encapsulation of metal atoms and clusters results in more complex redox behavior of EMFs since both the carbon cage and the endohedral cluster can exhibit redox activity. Especially interesting are endohedral (in cavea) electron transfer processes, in which the endohedral cluster is redox-active, whereas the carbon cage acts as an inert container transparent to electrons [8,9•]. An obvious prerequisite for the endohedral redox activity in EMF molecules is a localization of frontier molecular orbitals (HOMO or LUMO) on endohedral species. Experimentally, the endohedral redox processes can be identified via unexpected redox behavior (e.g., shifted potential when compared to analogous molecules) or via spectroscopic characterization of the charged species. Electron paramagnetic resonance (EPR) spectroscopy is an especially powerful tool, since EMFs with endohedral redox activity often exhibit rich hyperfine structure with large coupling constants in their ion radicals [10•].
This review is focused on the electrochemistry of EMFs featuring redox-active metal–metal bonds, and in particular on dimetallofullerenes (di-EMFs hereafter). First, we describe the electronic structure of di-EMFs from the molecular orbital (MO) point of view. This description forms a basis for the understanding of the redox behavior of three types of di-EMF: di-EMFs without metal–metal bonds, but with metal-based LUMO; di-EMFs with two-electron metal–metal bonds; and di-EMF with single-electron metal–metal bonds. Discussion of electrochemical properties of di-EMF is accompanied by the results of EPR spectroscopic measurements of their radical species.
Metal–metal bonding in dimetallofullerenes: theoretical description
Computational studies of di-EMF with Sc, Y, or lanthanides (metal is designated as M hereafter) show that these molecules feature metal–metal bonding molecular orbital, whose energy is close to the energy of the frontier cage-based MOs [11,12]. Whether the M–M bonding MO in a given di-EMF is the HOMO or the LUMO depends on the energy match between the metal-based and fullerene-based orbitals.
Figure 1a shows MO energy levels of two fullerene cages typical for di-EMFs, C80-Ih(7) and C82-C3v(8) (fullerene isomers are designated by their point group symmetry and the number in accordance with Fowler–Manolopoulos spiral algorithm [13]). Characteristic feature of C80-Ih(7) is the 4-fold degenerate orbital occupied by only two electrons. Jahn–Teller distortion reduces the symmetry and introduces a small gap between the HOMO and the 3-fold degenerate LUMO. The electronic structure of the molecule is very unstable, and C80-Ih(7) has never been obtained as an empty fullerene. However, if the LUMO is filled with six electrons, a stable structure with large band gap is obtained [14]. C80-Ih(7) is thus an archetypical cage for EMFs with 6-fold electron transfer from endohedral species to the fullerene [3].
Figure 1.
(a) Molecular orbital energy level of empty fullerene C80-Ih(7) and C82-C3v(8) compared to those of the metal dimers La2 and Lu2 (DFT calculations at the PBE/TZ2P level). Occupied MO levels of fullerenes are shown as black lines, unoccupied levels–as pink lines. Gray arrows indicate donation of six or four electrons from metal dimer to fullerene in corresponding dimetallofullerenes. (b) Frontier molecular orbitals (HOMO and LUMO) of La2@C80-Ih(7) and Lu2@C82-C3v(8).
C82-C3v(8) has small HOMO–LUMO gap, two low-lying unoccupied MOs, and a significant gap between the LUMO+1 and LUMO+2. The electronic structure of this fullerene is stabilized by addition of four electrons [15]. C82-C3v(8) (along with C82-Cs(6), which has similar electronic structure) is therefore the most abundant fullerene cage for EMFs with 4-fold electron transfer.
Also shown in Figure 1 are the energy levels of the occupied valence MOs in the two lanthanide dimers, La2 and Lu2. La2 has closed-shell electronic structure with six electrons occupying three MOs (hence (6s)σg2(5d)πu4 configuration) [16]. The energies of these MOs are considerably higher than the energy of the LUMO in C80-Ih(7), so when the La2 dimer is encapsulated inside this cage, a complete transfer of all six valence electrons to the fullerene occurs. The formal charge distribution in the resulting di-EMF molecule is then (La3+)2@C806−, the HOMO is localized on the fullerene, whereas the LUMO resembles the (6s)σg2 orbital of the pristine La2 dimer (Figure 1b). Thus, there is no La–La bonding in the non-charged La2@C80, but the LUMO of the molecule has the La–La bonding character, and the bond between metal atoms can be formed if the LUMO is populated by a surplus electron.
The lanthanide contraction results in a substantially different electronic structure of Lu2 when compared to that of La2. The ground state of Lu2 is a triplet, (6s)σg2(6s)σu2(5d)πu2 [16], with a significant splitting of the spin-up and spin-down orbitals (Figure 1a). These orbitals span a broader energy range than in La2. In particular, the (6s)σg2 level in Lu2 is ca. 2 eV lower in energy than in La2 and, even more importantly, it has lower energy than the LUMO of C80-Ih(7). As a result, the hypothetical Lu2@C80-Ih(7) has an open-shell electronic structure with five electrons transferred from Lu2 to the C80-Ih cage [17•]. C82-C3v(8) is a more suitable host for the Lu2 dimer than C80-Ih(7). In Lu2@C82-C3v(8), four electrons from the (6s)σu2(5d)πu2 levels of Lu2 are donated to the fullerene cage, whereas the (6s)σg2 orbital of Lu2 remains occupied. The formal charge distribution in the di-EMF is then (Lu2+)2@C824−. The Lu–Lu bonding orbital resembling the (6s)σg2 MO of Lu2 is the HOMO of Lu2@C82, whereas the LUMO is localized on the fullerene cage (Figure 1b).
Redox-active metal–metal bonds in dimetallofullerenes
Dimetallofullerenes with the metal-based LUMO
Early lanthanides, such as La, Ce, and less studied Pr and Nd, form di-EMFs with the transfer of all six valence electrons to the carbon cage. In addition to the C80-Ih(7) cage, several other fullerenes can act as acceptors of six electrons: La and/or Ce di-EMFs were reported for C72-D2(10611) [18,19], C76-Cs(17490) [20], C78-D3h(5) [21,22], C80-D5h(6) [23], and C100-D5(450) [24]. In all these di-EMFs, the M–M bonding MO is the LUMO, and hence metal–metal bonds are expected to be formed in the anionic state(s).
Electrochemical studies of La2@C2n (2n = 72, 78, 80) showed that these di-EMFs exhibit 2–3 reversible single-electron reduction steps. The first reduction of La2@C80-Ih occurs at −0.31 V (all redox potentials discussed hereafter are measured in o-dichlorobenzene and are referred to the Fe(Cp)2+/0 redox couple) [25]. Likewise, the first reductions of La2@C72 (−0.68 V) [26], La2@C78 (−0.40 V) [22], and La2@C80-D5h (−0.36 V) [23] are significantly more positive than for the EMFs with fullerene-based reductions (usually more negative than −1 V [1]). The first reduction potentials of analogous Ce di-EMFs are cathodicaly shifted by 0.04–0.13 V versus isostructural La di-EMFs (Table 1) [21,23,27,28].
Table 1. Redox potentials of di-EMFs featuring the M–M bonding HOMO or LUMO in comparison to selected clusterfullerenesa.
| EMF | E+2/+1 | E+1/0 | E0/−1 | E−1/−2 | E−2/−3 | gapECb | Ref. |
|---|---|---|---|---|---|---|---|
| Metal-based LUMO | |||||||
| La2@C72-D2(10611) | 0.75 | 0.24 | −0.68 | −1.92 | − | 0.92 | [26] |
| Ce2@C72-D2(10611) | 0.82 | 0.18 | −0.81 | −1.86 | − | 0.99 | [27] |
| La2@C76-Cs(17490) | 0.65 | 0.21 | −0.63 | −1.83 | −2.40 | 0.84 | [20] |
| La2@C78-D3h(5) | 0.62 | 0.26 | −0.40 | −1.84 | −2.28 | 0.66 | [22] |
| Ce2@C78-D3h(5) | 0.79 | 0.25 | −0.52 | −1.86 | −2.23 | 0.77 | [21] |
| La2@C80-D5h(6) | 0.78 | 0.22 | −0.36 | −1.72 | − | 0.58 | [23] |
| Ce2@C80-D5h(6) | 0.66 | 0.20 | −0.40 | −1.76 | −2.16 | 0.60 | [23] |
| La2@C80-Ih(7) | 0.95 | 0.56 | −0.31 | −1.72 | − | 0.87 | [25] |
| Ce2@C80-Ih(7) | 0.95 | 0.57 | −0.39 | −1.71 | − | 0.96 | [28] |
| Metal-based HOMO | |||||||
| Er2@C82-Cs(6) | 0.65 | 0.02 | −1.01 | −1.31 | − | 1.03 | [32••] |
| Lu2@C82-Cs(6) | 0.74 | 0.34 | −1.00 | −1.32 | −1.77 | 1.34 | [32••] |
| Er2S@C82-Cs(6)c | − | 0.39 | −1.01 | −1.85 | −2.21 | 1.40 | [32••] |
| Sc2S@C82-Cs(6)c | 0.65 | 0.39 | −0.98 | −1.12 | −1.73 | 1.37 | [50] |
| Sc2@C82-C3v(8) | − | 0.02 | −1.16 | −1.53 | −1.73 | 1.18 | [32••] |
| ErSc@C82-C3v(8) | − | 0.08 | −1.11 | −1.49 | −1.72 | 1.19 | [32••] |
| Er2@C82-C3v(8) | − | 0.13 | −1.14 | −1.41 | −1.83 | 1.27 | [32••] |
| YLu@C82-C3v(8) | − | 0.23 | −1.13 | − | − | 1.36 | [32••] |
| Lu2@C82-C3v(8) | 0.95 | 0.50 | −1.16 | −1.46 | −1.77 | 1.66 | [32••] |
| Er2S@C82-C3v(8)c | 0.88 | 0.51 | −0.98 | −1.21 | −1.70 | 1.49 | [32••] |
| Sc2S@C82-C3v(8)c | 0.96 | 0.52 | −1.04 | −1.19 | −1.63 | 1.56 | [50] |
| Lu2@C86-C2v(9) | − | 0.31 | −1.01 | −1.34 | − | 1.35 | [31] |
| Sc2C2@C86-C2v(9)c | − | 0.47 | −0.84 | −1.11 | −1.63 | 1.31 | [51] |
| Single-electron M–M bonding MO | |||||||
| Gd2@C79N-Ih(7) | − | 0.51 | −0.96 | −1.98 | − | 1.45 | [43] |
| La2@C80-CH2Ph | − | 0.15 | −0.92 | −1.34 | −1.64 | 0.97 | [45••] |
| Y2@C80-CH2Ph | 0.98 | 0.52 | −0.52 | −1.29 | −1.60 | 1.04 | [42••] |
| Gd2@C80-CH2Ph | 0.52 | −0.86 | −1.35 | 1.38 | [47] | ||
| Tb2@C80-CH2Ph | 0.51 | −0.79 | −1.36 | −1.71 | 1.30 | [47] | |
| Dy2@C80-CH2Ph | 0.98 | 0.52 | −0.60 | −1.28 | −1.58 | 1.12 | [42••] |
| Ho2@C80-CH2Ph | 0.51 | −0.54 | −1.33 | 1.05 | [47] |
All potentials are measured in o-dichlorobenzene solution and referenced versus Fe(Cp)2+/0 redox pair; redox processes involving M–M bonding orbitals are highlighted in bold;
gapEC is defined as E+1/0 − E0/−1 ;
clusterfullerenes with cage-based first oxidation steps, listed here for comparison to di-EMFs with the same fullerene cages.
Besides the value of the first reduction potential and its metal-dependence, another indication of the metal-based reduction in La and Ce di-EMFs is the difference between the first and the second reduction potentials, which amounts to 1.23–1.44 V (Table 1). For a fullerene redox process based on the same MO, the difference between the first and second reduction steps is usually within 0.4–0.5 V. The metal-based redox process results in a much larger potential difference for the consequent redox steps, because these steps are either based on the M–M bonding MO (with a much higher on-site Coulomb interaction than in the fullerene) or affect different MOs (one metal-based, and one delocalized over the carbon cage). Thus, both the high potential of the first reduction step and the large gap between the first and the second reduction potentials point to the population of the M–M bonding MO and hence formation of the single-electron M–M bond at the first reduction step.
Formation of the single-occupied La–La bonding MO in the [La2@C80-Ih]•− anion radical is further confirmed by EPR spectroscopy [29]. The M–M bonding orbitals in di-EMFs have hybrid spd character with large s-contribution, and population of such MOs by a single electron is expected to give paramagnetic species with large metal-based hyperfine constants [10•,11]. Indeed, huge isotropic 139La coupling constant of 364 G was reported in the radical anion [La2@C80-Ih]•− [29,30].
Dimetallofullerenes with metal-based HOMO
Due to the lanthanide contraction, the metals close to the end of the lanthanide row exhibit more covalent character in their compounds. In di-EMFs, Er and Lu give away only two electrons to the fullerene cage (the formal charge of the fullerene cage is thus −4). The remaining metal-based valence electrons then form the M–M bond via the sigma-type spd-hybrid MO. The most abundantly produced di-EMFs with four-fold charged fullerene cages are the two isomers of C82, Cs(6) and C3v(8) [31–34]. Structural characterization was also reported for several other Lu di-EMFs, including Lu2@C76-Td(1) [35], Lu2@C84-D2d(23), and Lu2@C86-C2v(9) [31]. If the M–M bonding MO is the HOMO (as predicted by theory), these di-EMFs should feature a metal–metal bond already in the pristine non-charged state, and this bond should be electrochemically active in the oxidation processes.
The experimental confirmation of the metal–metal bonding in di-EMFs is not very straightforward. The formal charge of the fullerene cage in M2@C82 can be deduced from Vis-NIR spectroscopic measurements. UV-vis-NIR absorption spectra of the di-EMFs with the same fullerene cage and different metals are virtually identical. Presumably, the excitation originating from the metal-based HOMO have very low intensity and cannot be observed, resulting in the dominance of the π →π* transitions in the fullerene cages. Similar spectra are also observed for sulfide clusterfullerenes M2S@C82 [32••,36] or carbide clusterfullerenes M2C2@C82 [12,37]. In cluster-fullerenes, the non-metal endohedral entity bears a negative charge (C22−, O2−, or S2−), the metal atoms are in their 3+ state, whereas the cage has the negative charge of −4. Such clusterfullerenes do not feature metal–metal bonds, and their frontier MOs are usually localized on the fullerene cage [6•]. The close resemblance of the absorption spectra of di-EMF and clusterfullerenes proves that the carbon cage in these EMFs has the same formal charge, −4. Thus, +2 oxidation state of metal atoms appears natural. However, the presence of the M–M bond does not automatically follow from the oxidation state. X-ray absorption spectra at the M4,5 edge (3d→4f excitations commonly used in the studies of lanthanides) did not show substantial difference between Er2@C82 or Er2C2@C82 [38]. However, both Er2+ and Er3+ states in EMFs feature the same 4f12 occupation, and therefore absorption at the M4,5 edge may be not sensitive enough to the difference in the valence orbital populations.
Electrochemistry provides more straightforward approach to the problem. If the M–M bonding MO is indeed the redox-active HOMO of di-EMFs, the first oxidation potential should exhibit pronounced metal-dependence in contrast to the first reduction potential, which corresponds to the cage-based LUMO and is not expected to vary much from metal to metal. Indeed, electrochemistry reveals pronounced differences in the electronic structure of Er2@C82 and Lu2@C82 [32••]. Their reduction potentials are rather similar (Figure 2a; note that C82-C3v isomer exhibits irreversible reduction steps, whereas reduction steps of C82-Cs isomers are fully reversible). Such a similarity of the potentials points to the fullerene-based nature of the underlying redox steps, in agreement with DFT prediction. On the contrary, oxidation potentials of Er2@C82 and Lu2@C82 are strongly metal-dependent. Er2@C82 isomers have their first oxidation step at ca. 0.3–0.4 V lower potentials than Lu-counterparts (Table 1). The same trend was observed for M2@C82-C3v structures with other metals, including Sc2@C82 and mixed-metal ErSc@C82 and YLu@C82 di-EMFs [32••]. With almost identical reduction potential, they exhibit large variability of the first oxidation potentials (Figure 2b). Metal-dependence of the first oxidation potential in di-EMFs confirms the computationally predicted metal– metal bonding HOMO in these molecules (Figure 2c). Lu has the lowest energy of the M–M bonding HOMO, and hence Lu di-EMFs exhibit the highest oxidation potentials when compared to other metals. In fact, oxidation potentials of Lu-di-EMFs are close to the cage-based oxidation potentials of sulfide clusterfullerenes M2S@C82 (Table 1), and it is hard to distinguish if the first oxidation step of Lu2@C82 isomers is metal- or fullerene-based. Lower oxidation potentials of di-EMFs with other metals unequivocally point of the metal-based processes.
Figure 2.
(a) Cyclic voltammetry of Er2@C82 and Lu2@C82 dimetallofullerenes with C3v (8) and Cs(6) cage isomers in o-dichlorobenzene/TBAPF6 solution at 100 mV s−1; whereas the first reduction potentials of Lu2@C82 and Er2@C82 with the same fullerene cage are virtually identical (denoted by blue dashed line), the first oxidation potentials are different by more than 0.3 V (red lines); (b) square wave voltammetry of several M2@C82-C3 v(8) at the first oxidation step (M2 = Lu2, YLu, Er2, ErSc, Sc2); (c) HOMO orbitals for Lu2@C82, Y2@C82, and YLu@C82 ; (d) EPR spectrum of Sc2@C82+ cation in o-dichlorobenzene at room temperature, a(45Sc) = 199.2 G, g = 1.994; the lines show assignment of the peaks in terms of |I, mI) nuclear spin quantum numbers of the Sc2 dimer. Reproduced with permission from the Ref. [32••].
Single-electron oxidation of di-EMF with M–M bonding HOMO produces a single-occupied metal-based orbital with unprecedented spin properties. Large contribution of metal s-atomic orbital to the M–M HOMO of M2@C82 yields a large isotropic hyperfine coupling constant for metals with non-zero nuclear spin in [M2@C82]•+ cation radicals. A striking example is the cation radical of Sc2@C82, which at room temperature in o-dichlorobenzene solution exhibits well-resolved EPR spectrum with the hyperfine structure spanning 2800 G (Figure 2d). Instead of 15 lines expected for two equivalent Sc with nuclear spin of 7/2, experimental spectrum comprises 64 lines caused by additional splitting due to the large 45Sc hyperfine constant, a(45Sc) = 199.2 G [32••]. Formation of the single-electron Er–Er bond in [Er2@C82-C3v]•+ was supported by SQUID magnetometry. The oxidation of Er2@C82 strongly modified the spin state of the endohedral Er2 unit, presumably creating a three-center [Er3+–e–Er3+] system with stronger exchange interactions than in the pristine Er2@C82 [32••].
Dimetallofullerenes with single-electron metal–metal bond
Whereas early and late lanthanides tends to form di-EMFs with tri- and di-valent state of metals, respectively, yttrium and lanthanides in the middle part of the lanthanide row (Gd–Ho) give di-EMF with even more peculiar electronic structure. Computational studies of M2@C80-Ih (M = Y, Lu) showed that the ground electronic state for these di-EMF is a triplet [17•]. The M–M bonding MO is occupied by a single electron, and another unpaired spin is delocalized over the fullerene cage. The formal charge distribution is then (M2)5+@C805−. During the extraction of fullerenes from the arc-discharge soot by standard fullerene solvents (such as CS2 or toluene) these molecules remain insoluble, presumably due to polymerization or aggregation with the soot particles.
Electronic structure of such M2@C80-Ih di-EMFs can be stabilized by addition of an electron, which yields to closed-shell electronic structure of the fullerene cage, (M2)5+@C806− [39]. Indeed, the synthesis of M2@C80 derivatives was accomplished when EMFs were extracted from the soot with N,N-dimethylformamide, which is known to form fullerene anions during extraction (Figure 3a) [40,41]. Chemical derivatization with a single radical group R (R = CF3 or benzyl CH2Ph) is another way to quench the cage-based radical in M2@C80-Ih [17•,42••]. Finally, addition of an electron is equivalent to substitution of one carbon atom by nitrogen. C79N5− is isoelectronic to C806−, and stable M2@C79N compounds were obtained in Dorn’s group for M = Y, Gd, and Tb [43,44]. The common feature of M2@C80−, M2@C79N, or M2@C80(R) is the single-electron M–M bond stabilized inside the fullerene. For M = Y, localization of the spin density on the Y–Y bonding MO can be confirmed by EPR spectroscopy, which revealed similar spectra in all three types of EMFs with large isotropic 89Y hyperfine coupling constants near of 80 G (Figure 3b,c) [17•,42••,44]. Formation of the single-electron La–La bond in La2@C80(CH2Ph) and similar radical monoadducts of La2@C80-Ih was also confirmed by EPR spectroscopy and single-crystal X-ray diffraction [45••,46].
Figure 3.
(a) Schematic description of the electron distribution between M–M bonding MO and fullerene cage in dimetallofullerenes M2@C80-Ih (M = Y, Tb, Dy, etc.) and a chemical route to stabilize these structures via reduction and subsequent nucleophilic substitution yielding air-stable M2@C80(CH2Ph) monoadduct. (b) EPR spectra of the toluene solution of Y2@C80(CH2Ph) at room temperature and at 150 K (below the freezing point of the solvent); the isotropic RT spectrum has g-factor of 1.9733 and the aiso(89Y) value of 223.8 MHz; the axial spectral pattern in frozen solution is reproduced by g⊥ = 1.9620, gǁ = 1.9982, a⊥(89Y) = 208.0 MHz, aǁ(89Y) = 245.9 MHz; (c) spin density distribution in Y2@C80(CH2Ph) computed at the PBE0/TZVP level; (d) square wave voltammetry of M2@C80(CH2Ph) (M = Y, Dy, and Tb), black vertical bars denote redox potentials of La2@C80(CH2Ph) from Ref. [45••], dotted lines denote the first oxidation (cyan) and the first reduction (red) potentials. Based on the data from Ref. [42••].
Potentially, the half-occupied M–M bonding orbital can be redox active both in reduction and oxidation processes. However, DFT calculations predict large energy difference between occupied and unoccupied counterparts of the MO [42••]. As a result, the occupied component of the M–M bonding MO in La2@C80(CH2Ph) is predicted to be the HOMO, whereas the LUMO is localized on the fullerene cage. In Y2@C80(CH2Ph), analogous calculations showed metal-based LUMO and fullerene-based HOMO.
Y and Gd–Ho M2@C80(CH2Ph) derivatives exhibit virtually identical oxidation potential at +0.51–0.52 V (Table 1, Figure 3d) [32••,47]. The lack of the metal dependence is an indication of the fullerene-based oxidation in these di-EMFs, in agreement with DFT prediction for Y2@C80(CH2Ph). With the first oxidation potential at +0.15 V [45••], La2@C80(CH2Ph) is an obvious outlier exhibiting the metal-based oxidation. Thus, the metal–metal bonding MO of La2@C80(CH2Ph) is depopulated in the oxidation process, whereas for other metals the single-electron M–M bond is not affected.
The first reduction potentials of M2@C80(CH2Ph) derivatives are different and span the range from −0.52 V in Y2@C80(CH2Ph) to −0.92 V in La2@C80(CH2Ph). In the Gd–Ho row, the potential is changing gradually with the size of the lanthanide, more negative values corresponding to larger ionic radii. This behavior is consistent with the metal-based reduction for Y and medium-size lanthanides. Hence, the single-electron M–M bond turns to a two-electron bond in their monoanions. In La2@C80(CH2Ph), the process switches to the fullerene-based reduction.
Electrochemical studies of other di-EMFs with single-electron M–M bond were reported so far only for Gd2@C79N [43]. Its first oxidation potential at +0.51 V is very close to that of M2@C80(CH2Ph) derivatives with fullerene-based oxidation (Table 1). The first reduction at −0.96 V is more negative than in any M2@C80(CH2Ph), including La2@C80(CH2Ph). Computational studies showed that Gd2@C79N has two low-energy unoccupied MOs, one Gd-based and one delocalized over the fullerene, thus making it hard to distinguish between the fullerene- and metal-based reductions. However, the large difference between the first and the second reduction potentials of almost 1 V (Table 1) indicates that the first reduction of Gd2@C79N may indeed involve the Gd–Gd bonding MO.
Outlook
The unique environment of endohedral fullerene provides a possibility to stabilize exotic species with unconventional bonding situation, such as lanthanide dimers with metal–metal bonds. Whereas many lanthanide complexes with low oxidation states have been synthesized [48••], no other molecular compounds with lanthanide–lanthanide bonds have been reported so far [49••]. Furthermore, the M–M bonding MOs in dimetallofullerenes are redox active and undergo single-electron reduction or oxidation, which leads to radical species with single-electron M–M bonding MOs. Electrochemistry is thus found to be a convenient technique to study metal–metal bonds in fullerenes. Redox variability of the population of the lanthanide–lanthanide bonding MOs in di-EMFs is very useful for tuning their magnetic properties. The presence of the unpaired valence electron in lanthanide-based di-EMFs results in giant exchange interactions and coupling of local 4f-derived spins and unpaired spin in the M–M bonding MO into a larger “superspin”. If lanthanides with large magnetic anisotropy (such as Dy or Tb) are coupled this way, single molecule magnets with high blocking temperature of magnetization can be obtained [42••]. Semi-occupied M–M bonding MO is also essential for the spin-polarized electronic transport through single fullerene molecules, which can lead to single-molecule electronic and spintronic devices.
Acknowledgments
We acknowledge European Research Council (ERC) under the European Union Horizon 2020 research and innovation programme (grant agreement no 648295 “GraM3”) and Deutsche Forschungsgemeinschaft (grant PO 1602/4-1).
Abbreviations
- EMF
endohedral metallofullerene
- di-EMF
dimetallofullerene
- MO
molecular orbital
- HOMO
highest occupied molecular orbital
- LUMO
lowest unoccupied molecular orbital
- EPR
electron paramagnetic resonance
- M–M bond
metal–metal bond
- Vis-NIR
visible and near-infrared
- DFT
density functional theory.
References and recommended reading
Papers of particular interest, published within the period of review, have been highlighted as:
• Paper of special interest.
•• Paper of outstanding interest.
- 1.Popov AA, Yang S, Dunsch L. Endohedral fullerenes. Chem Rev. 2013;113:5989–6113. doi: 10.1021/cr300297r. [DOI] [PubMed] [Google Scholar]
- 2.Lu X, Feng L, Akasaka T, Nagase S. Current status and future developments of endohedral metallofullerenes. Chem Soc Rev. 2012;41:7723–7760. doi: 10.1039/c2cs35214a. [DOI] [PubMed] [Google Scholar]
- 3.Wang T, Wang C. Endohedral metallofullerenes based on spherical Ih-C80 cage: molecular structures and paramagnetic properties. Acc Chem Res. 2014;47:450–458. doi: 10.1021/ar400156z. [DOI] [PubMed] [Google Scholar]
- 4.Rodriguez-Fortea A, Balch AL, Poblet JM. Endohedral metallofullerenes: a unique host-guest association. Chem Soc Rev. 2011;40:3551–3563. doi: 10.1039/c0cs00225a. [DOI] [PubMed] [Google Scholar]
- 5.Yang S, Wei T, Jin F. When metal clusters meet carbon cages: endohedral clusterfullerenes. Chem Soc Rev. 2017;46:5005–5058. doi: 10.1039/c6cs00498a. [DOI] [PubMed] [Google Scholar]
- 6.Popov AA. Electrochemistry and frontier molecular orbitals of endohedral metallofullerenes. In: Popov AA, editor. Endohedral Fullerenes: Electron Transfer and Spin. Cham: Springer International Publishing; 2017. pp. 35–62. [• The most recent overview of electrochemical properties of endohedral fullerenes, accompanied by the analysis of their molecular orbitals] [Google Scholar]
- 7.Echegoyen L, Echegoyen LE. Electrochemistry of fullerenes and their derivatives. Acc Chem Res. 1998;31:593–601. [Google Scholar]
- 8.Popov AA, Dunsch L. Electrochemistry in cavea: endohedral redox reactions of encaged species in fullerenes. J Phys Chem Lett. 2011;2:786–794. [Google Scholar]
- 9.Zhang Y, Popov AA. Transition-metal and rare-earth-metal redox couples inside carbon cages: fullerenes acting as innocent ligands. Organometallics. 2014;33:4537–4549. [• An overview of EMFs exhibiting endohedral electrochemical processes, i.e. featuring redox activity of endohedral clusters] [Google Scholar]
- 10.Popov AA. Ion radicals of endohedral metallofullerenes studied by EPR spectroscopy. In: Popov AA, editor. Endohedral Fullerenes: Electron Transfer and Spin. Cham: Springer International Publishing; 2017. pp. 183–198. [• An overview of electron paramagnetic resonance studies of ion radicals of endohedral metallofullerenes] [Google Scholar]
- 11.Popov AA, Avdoshenko SM, Pendás AM, 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. doi: 10.1039/c2cc32568c. [DOI] [PubMed] [Google Scholar]
- 12.Kurihara H, Lu X, Iiduka Y, Mizorogi N, Slanina Z, Tsuchiya T, Nagase S, Akasaka T. Sc2@C3v(8)-C82 vs. Sc2C2@C3v(8)-C82: drastic effect of C2 capture on the redox properties of scandium metallofullerenes. Chem Commun. 2012;48:1290–1292. doi: 10.1039/c2cc16422a. [DOI] [PubMed] [Google Scholar]
- 13.Fowler P, Manolopoulos DE. An Atlas of Fullerenes. Oxford, U.K.: Clarendon Press; 1995. [Google Scholar]
- 14.Campanera JM, Bo C, Poblet JM. General rule for the stabilization of fullerene cages encapsulating trimetallic nitride templates. Angew Chem Int Ed. 2005;44:7230–7233. doi: 10.1002/anie.200501791. [DOI] [PubMed] [Google Scholar]
- 15.Valencia R, Rodríguez-Fortea A, Poblet JM. Understanding the stabilization of metal carbide endohedral fullerenes M2C2@C82 and related systems. J Phys Chem A. 2008;112:4550–4555. doi: 10.1021/jp800419d. [DOI] [PubMed] [Google Scholar]
- 16.Cao X, Dolg M. Pseudopotential study of lanthanum and lutetium dimers. Theor Chem Acc. 2002;108:143–149. [Google Scholar]
- 17.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. [• Computational study of M2@C80 molecules, which revealed unusual triplet ground state with single electron on metal–metal bonding orbital for Y2@C80 and Lu2@C80] [Google Scholar]
- 18.Zhao Y-L, Yu H-T, Lian Y-F. Experimental and theoretical evaluation of structures of Pr2@C72 and its functionalized adduct with adamantylidene carbene. RSC Adv. 2016;6:115113–115119. [Google Scholar]
- 19.Stevenson S, Burbank P, Harich K, Sun Z, Dorn HC, van Loosdrecht PHM, deVries MS, Salem JR, Kiang CH, Johnson RD, Bethune DS. La2@C72: metal-mediated stabilization of a carbon cage. J Phys Chem A. 1998;102:2833–2837. [Google Scholar]
- 20.Suzuki M, Mizorogi N, Yang T, Uhlik F, Slanina Z, Zhao X, Yamada M, Maeda Y, Hasegawa T, Nagase S, et al. La2@Cs(17 490)-C76: a new non-IPR dimetallic metallofullerene featuring unexpectedly weak metal–pentalene interactions. Chem Eur J. 2013;19:17125–17130. doi: 10.1002/chem.201302821. [DOI] [PubMed] [Google Scholar]
- 21.Yamada M, Wakahara T, Tsuchiya T, Maeda Y, Kako M, Akasaka T, Yoza K, Horn E, Mizorogi N, Nagase S. Location of the metal atoms in Ce2@C78 and its bis-silylated derivative. Chem Commun. 2008:558–560. doi: 10.1039/b712568b. [DOI] [PubMed] [Google Scholar]
- 22.Cao BP, Wakahara T, Tsuchiya T, Kondo M, Maeda Y, Rahman GMA, Akasaka T, Kobayashi K, Nagase S, Yamamoto K. Isolation, characterization, and theoretical study of La2@C78. J Am Chem Soc. 2004;126:9164–9165. doi: 10.1021/ja048599g. [DOI] [PubMed] [Google Scholar]
- 23.Yamada M, Mizorogi N, Tsuchiya T, Akasaka T, Nagase S. Synthesis and characterization of the D5h isomer of the endohedral dimetallofullerene Ce2@C80: two-dimensional circulation of encapsulated metal atoms inside a fullerene cage. Chem Eur J. 2009;15:9486–9493. doi: 10.1002/chem.200900713. [DOI] [PubMed] [Google Scholar]
- 24.Beavers CM, Jin H, Yang H, Wang Z, Wang X, Ge H, Liu Z, Mercado BQ, Olmstead MM, Balch AL. Very large, soluble endohedral fullerenes in the series La2C90 to La2C138: isolation and crystallographic characterization of La2@D5(450)-C100. J Am Chem Soc. 2011;133:15338–15341. doi: 10.1021/ja207090e. [DOI] [PubMed] [Google Scholar]
- 25.Suzuki T, Maruyama Y, Kato T, Kikuchi K, Nakao Y, Achiba Y, Kobayashi K, Nagase S. Electrochemistry and ab-initio study of the dimetallofullerene La2@C80. Angew Chem Int Ed. 1995;34:1094–1096. [Google Scholar]
- 26.Lu X, Nikawa H, Nakahodo T, Tsuchiya T, Ishitsuka MO, Maeda Y, Akasaka T, Toki M, Sawa H, Slanina Z, et al. Chemical understanding of a non-IPR metallofullerene: stabilization of encaged metals on fused-pentagon bonds in La2@C72. J Am Chem Soc. 2008;130:9129–9136. doi: 10.1021/ja8019577. [DOI] [PubMed] [Google Scholar]
- 27.Yamada M, Wakahara T, Tsuchiya T, Maeda Y, Akasaka T, Mizorogi N, Nagase S. Spectroscopic and theoretical study of endohedral dimetallofullerene having a non-IPR fullerene cage: Ce2@C72. J Phys Chem A. 2008;112:7627–7631. doi: 10.1021/jp804260d. [DOI] [PubMed] [Google Scholar]
- 28.Yamada M, Nakahodo T, Wakahara T, Tsuchiya T, Maeda Y, Akasaka T, Kako M, Yoza K, Horn E, Mizorogi N, et al. Positional control of encapsulated atoms inside a fullerene cage by exohedral addition. J Am Chem Soc. 2005;127:14570–14571. doi: 10.1021/ja054346r. [DOI] [PubMed] [Google Scholar]
- 29.Kato T. Metal dimer and trimer within spherical carbon cage. J Mol Struct. 2007;838:84–88. [Google Scholar]
- 30.Tsuchiya T, Wielopolski M, Sakuma N, Mizorogi N, Akasaka T, Kato T, Guldi DM, Nagase S. Stable radical anions inside fullerene cages: formation of reversible electron transfer systems. J Am Chem Soc. 2011;133:13280–13283. doi: 10.1021/ja205391v. [DOI] [PubMed] [Google Scholar]
- 31.Shen W, Bao L, Wu Y, Pan C, Zhao S, Fang H, Xie Y, Jin P, Peng P, Li F-F, Lu X. Lu2@C2n (2n = 82, 84, 86) crystallographic evidence of direct Lu–Lu bonding between two divalent lutetium ions inside fullerene cages. J Am Chem Soc. 2017;139:9979–9984. doi: 10.1021/jacs.7b04421. [Synthesis of Lu-dimetallofullerenes, and their single-crystal X-ray diffraction studies showing the short Lu–Lu distances.] [DOI] [PubMed] [Google Scholar]
- 32.Samoylova NA, Avdoshenko SM, Krylov DS, Thompson HR, Kirkhorn A, Rosenkranz M, Schiemenz S, Ziegs F, Wolter AUB, Yang S, et al. Confining the spin between two metal atoms within the carbon cage: redox-active metal-metal bonds in dimetallofullerenes and their stable cation radicals. Nanoscale. 2017;9:7977–7990. doi: 10.1039/c7nr02288c. [••Extended electrochemical studies of M2@C82 dimetallofullerenes proving the presence of metal–metal bonds and their single-electron oxidation.] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Olmstead MM, Lee HM, Stevenson S, Dorn HC, Balch AL. Crystallographic characterization of Isomer 2 of Er2@C82 and comparison with Isomer 1 of Er2@C82. Chem Commun. 2002:2688–2689. doi: 10.1039/b209270k. [DOI] [PubMed] [Google Scholar]
- 34.Olmstead MM, de Bettencourt-Dias A, Stevenson S, Dorn HC, Balch AL. Crystallographic characterization of the structure of the endohedral fullerene {Er2@C82 Isomer I} with Cs cage symmetry and multiple sites for erbium along a band of ten contiguous hexagons. J Am Chem Soc. 2002;124:4172–4173. doi: 10.1021/ja0116019. [DOI] [PubMed] [Google Scholar]
- 35.Umemoto H, Ohashi K, Inoue T, Fukui N, Sugai T, Shinohara H. Synthesis and UHV-STM observation of the Td-symmetric Lu metallofullerene: Lu2@C76(Td) Chem Commun. 2010;46:5653–5655. doi: 10.1039/c0cc00824a. [DOI] [PubMed] [Google Scholar]
- 36.Chen C-H, Krylov DS, Avdoshenko SM, Liu F, Spree L, Yadav R, Alvertis A, Hozoi L, Nenkov K, et al. Selective arc-discharge synthesis of Dy2S-clusterfullerenes and their isomer-dependent single molecule magnetism. Chem Sci. 2017;8:6451–6465. doi: 10.1039/c7sc02395b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ito Y, Okazaki T, Okubo S, Akachi M, Ohno Y, Mizutani T, Nakamura T, Kitaura R, Sugai T, Shinohara H. Enhanced 1520nm photoluminescence from Er3+ ions in di-erbium-carbide metallofullerenes (Er2C2)@C82 (isomers I, II and III) ACS Nano. 2007;1:456–462. doi: 10.1021/nn700235z. [DOI] [PubMed] [Google Scholar]
- 38.Okimoto H, Kitaura R, Nakamura T, Ito Y, Kitamura Y, Akachi T, Ogawa D, Imazu N, Kato Y, Asada Y, et al. Element-specific magnetic properties of di-erbium Er2@C82 and Er2C2@C82 metallofullerenes: a synchrotron soft X-ray magnetic circular dichroism study. J Phys Chem C. 2008;112:6103–6109. [Google Scholar]
- 39.Velloth A, Imamura Y, Kodama T, Hada M. Theoretical Insights into the electronic structures and stability of dimetallofullerenes M2@Ih-C80. J Phys Chem C. 2017;121:18169–18177. [Google Scholar]
- 40.Kareev IE, Bubnov VP, Yagubskii EB. Endohedral gadolinium-containing metallofullerenes in the trifluoromethylation reaction. Russ Chem Bull. 2008;57:1486–1491. [Google Scholar]
- 41.Kareev IE, Lebedkin SF, Bubnov VP, Yagubskii EB, Ioffe IN, Khavrel PA, Kuvychko IV, Strauss SH, Boltalina OV. Trifluoromethylated endohedral metallofullerenes: synthesis and characterization of Y@C82(CF3)5. Angew Chem Int Ed. 2005;44:1846–1849. doi: 10.1002/anie.200461497. [DOI] [PubMed] [Google Scholar]
- 42.Liu F, Krylov DS, Spree L, Avdoshenko SM, Samoylova NA, Rosenkranz M, Kostanyan A, Greber T, Wolter AUB, Büchner B, Popov AA. Single molecule magnet with an unpaired electron trapped between two lanthanide ions inside a fullerene. Nat Commun. 2017;8:16098. doi: 10.1038/ncomms16098. [••Stabilization of M2@C80 (M = Y, Dy) di-EMFs with single-electron metal–metal bonds in the form of air-stable benzyl monoadducts, their EPR, electrochemical, and magnetic properties] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Fu W, Zhang J, Fuhrer T, Champion H, Furukawa K, Kato T, Mahaney JE, Burke BG, Williams KA, Walker K, et al. Gd2@C79N: isolation, characterization, and monoadduct formation of a very stable heterofullerene with a magnetic spin state of S = 15/2. J Am Chem Soc. 2011;133:9741–9750. doi: 10.1021/ja202011u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Zuo T, Xu L, Beavers CM, Olmstead MM, Fu W, Crawford TD, Balch AL, Dorn HC. 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. doi: 10.1021/ja802417d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Bao L, Chen M, Pan C, Yamaguchi T, Kato T, Olmstead MM, Balch AL, Akasaka T, Lu X. Crystallographic evidence for direct metal–metal bonding in a stable open-shell La2@Ih-C80 derivative. Angew Chem Int Ed. 2016;55:4242–4246. doi: 10.1002/anie.201511930. [• Photochemical synthesis of La2@C80(CH2Ph) with single-electron La–La bond; its structural, EPR, and electrochemical studies.] [DOI] [PubMed] [Google Scholar]
- 46.Yamada M, Kurihara H, Suzuki M, Saito M, Slanina Z, Uhlik F, Aizawa T, Kato T, Olmstead MM, et al. Hiding and recovering electrons in a dimetallic endohedral fullerene: air-stable products from radical additions. J Am Chem Soc. 2015;137:232–238. doi: 10.1021/ja509956y. [DOI] [PubMed] [Google Scholar]
- 47.Samoylova N.A., Liu F., Spree L., Popov A.A.: unpublished results. 2017.
- 48.Evans WJ. Tutorial on the role of cyclopentadienyl ligands in the discovery of molecular complexes of the rare-earth and actinide metals in new oxidation states. Organometallics. 2016;35:3088–3100. [•• The tutorial includes overview of Lanthanides with cyclopentadienyl ligands featuring divalent oxidation state.] [Google Scholar]
- 49.Liddle ST. Molecular Metal–Metal Bonds. Wiley-VCH Verlag GmbH & Co. KGaA; 2015. [•• Comprehensive overview of metal–metal bonding in molecular compounds.] [Google Scholar]
- 50.Mercado BQ, Chen N, Rodriguez-Fortea A, Mackey MA, Stevenson S, Echegoyen L, Poblet JM, Olmstead MM, Balch AL. The shape of the Sc2(μ2-S) unit trapped in C82: crystallographic, computational, and electrochemical studies of the isomers, Sc2(μ2-S)@Cs(6)-C82 and Sc2(μ2-S)@C3v(8)-C82. J Am Chem Soc. 2011;133:6752–6760. doi: 10.1021/ja200289w. [DOI] [PubMed] [Google Scholar]
- 51.Chen C-H, Ghiassi KB, Cerón MR, Guerrero-Ayala MA, Echegoyen L, Olmstead MM, Balch AL. Beyond the butterfly: Sc2C2@C2v(9)-C86, an endohedral fullerene containing a planar, twisted Sc2C2 unit with remarkable crystalline order in an unprecedented carbon cage. J Am Chem Soc. 2015;137:10116–10119. doi: 10.1021/jacs.5b06425. [DOI] [PubMed] [Google Scholar]



