Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Unwilling U–U bonding in U2@C80: cage-driven metal–metal bonds in di-uranium fullerenes

Cina Foroutan-Nejad ab, Jan Vícha bc, Radek Marek b, Michael Patzschke d and Michal Straka *a
aInstitute of Organic Chemistry and Biochemistry, Academy of Sciences, Flemingovo nám. 2., CZ-16610, Prague, Czech Republic. E-mail: straka@uochb.cas.cz
bCEITEC - Central European Institute of Technology, Masaryk University, Kamenice 5/A4, CZ-62500 Brno, Czech Republic
cCentre of Polymer Systems, University Institute, Tomas Bata University in Zlin, Trida T. Bati, 5678, CZ-76001, Zlin, Czech Republic
dHelmholtz-Zentrum Dresden-Rossendorf, POB 510119, DE-01314, Dresden, Germany

Received 22nd July 2015 , Accepted 21st August 2015

First published on 24th August 2015


Abstract

Endohedral actinide fullerenes are rare and a little is known about their molecular properties. Here we characterize the U2@C80 system, which was recently detected experimentally by means of mass spectrometry (Akiyama et al., JACS, 2001, 123, 181). Theoretical calculations predict a stable endohedral system, 7U2@C80, derived from the C80:7 IPR fullerene cage, with six unpaired electrons. Bonding analysis reveals a double ferromagnetic (one-electron-two-center) U–U bond at an rU–U distance of 3.9 Å. This bonding is realized mainly via U(5f) orbitals. The U–U interaction inside the cage is estimated to be about −18 kcal mol−1. U–U bonding is further studied along the U2@Cn (n = 60, 70, 80, 84, 90) series and the U–U bonds are also identified in U2@C70 and U2@C84 systems at rU–U ∼ 4 Å. It is found that the character of U–U bonding depends on the U–U distance, which is dictated by the cage type. A concept of unwilling metal–metal bonding is suggested: uranium atoms are strongly bound to the cage and carry a positive charge. Pushing the U(5f) electron density into the U–U bonding region reduces electrostatic repulsion between enclosed atoms, thus forcing U–U bonds.


1 Introduction

Endohedral fullerenes and particularly endohedral metallofullerenes (EMF) have been extensively studied as promising materials for practical applications. Despite the great developments in EMF science1–13 little is known of actinide endohedral fullerenes and their properties. Most of the experimentally reported actinide EMFs to date were only observed as signals in the time-of-flight mass spectra (TOF-MS), as for example U@C2n (n = 14–36) and some of the U2@C2n (n = 25–30) systems.14,15 Akiyama et al. reported a series of An@C82 (An = U, Np, Am) as well as An@C84 (An = Th, Pa) fullerenes,16–19 and also U2@C80 that is of interest in this work. U@C82 and Th@C84 were prepared in larger amounts to be characterized by UV/vis/NIR spectroscopy.16–19 The experimental formation of U@C28 was studied recently.20

Actinide fullerenes have also attracted the attention of theoreticians. Mainly the An@C28 compounds were studied.21–26 The An@C26 and An@C40 series27,28 and related compounds, such as U@C36, Pu@C24, and U@C82, were investigated, too.28–31 To the best of our knowledge, the experimentally observed U2@C80 molecule17 has not been studied yet, and is the main concern of the present work.

The presence of two actinide atoms in a fullerene cage brings another interesting aspect that makes the endohedral actinide fullerenes attractive – the possibility of forming actinide–actinide bonds in the interior of a fullerene. Although numerous examples of metal–metal bonds for d-block elements have been documented in the transition-metal chemistry, actinide–actinide bonds are rare. The question of the existence of actinide–actinide bonding dates back to the early studies by Cotton et al.32 and was revived by Gagliardi and Roos in 2005 in a study on U2 system,33 which is experimentally known,34 followed by sequels on actinide diatomics,35,36 and studies of various compounds with actinide–actinide bonds.37–43

Endohedral U–U bonding was suggested in 2007 by Wu and Lu44 who studied theoretically the U2@C60 system, observed previously in TOF-MS experiments.14,15 It was found, based on the MO framework, that the two U atoms confined in C60 form six one-electron-two-center (1e-2c, or ferromagnetic) metal–metal bonds at a calculated minimum U–U distance, rU–U = 2.72 Å. Infante et al.45 argued that the multiple U–U bonding in U2@C60 is, in fact, forced by the small interior of the cage. Hypothetical U2@C70 and U2@C84 fullerenes were calculated therein45 but the U–U bonding in these systems was not investigated, possibly because of the calculated large U–U separation, rU–U ∼ 3.9 Å. Dai et al. predicted that in hypothetical U2@C90, the uranium atoms separate to rU–U ∼ 6.1 Å.46

A recent study has predicted the UGd@C60 analogue of U2@C60 fullerene to have a large encapsulation energy and a high-spin 11-et ground state with a twofold one-electron U–Gd bond.47 Studies of U2@C61 revealed that the exohedral carbon atom has a strong influence on the U–U distance and ground-state spin multiplicity. Such defects can be used for tuning the electronic properties of EMFs.47,48

Endohedral metal–metal bonding has been recently discussed in some experimentally known lanthanide and transition-metal fullerenes, for example, in Y2@C79N,49 Lu2@C76,50–52 and anionic La2@C80 fullerenes.53 For more examples and references, see ref. 53 by Popov et al., where the topic of endohedral metal–metal bonding is reviewed and studied in detail.

In this work we characterize fullerene U2@C80 by means of theoretical calculations. A stable endohedral system with large encapsulation energy for U2 in the C80 cage is found. The energy and bonding analysis of U2@C80 provides evidence for metal–metal bonding interactions between the trapped uranium atoms. To further reveal the general trends in the endohedral U–U bonding we investigate a series of U2@Cn (n = 60, 70, 80, 84, 90) fullerenes and show newly the evidence for U–U bonding in hypothetical U2@C70 and U2@C84 cages as well as a correlation between the character of the U–U bonding and U–U distance inside a fullerene cage.

2 Methods

2.1 Molecular structure and properties

Following the previous experience,45 calculations were done with the BP86 density functional using the def-SVP basis sets for C and U atoms54,55 as implemented in the Turbomole 6.3.1 code. This basis set includes the effective core potential (SDD) for uranium atoms.55 The structure search was done using Turbomole 6.3.1 and Gaussian 09 programs.56,57 Natural population analysis (NPA) as implemented in the natural bond orbital58,59 analysis implemented in Gaussian 09 was used. Molecular Orbitals were plotted using the Multifwn60 and VMD software.61

The search for the geometry of the U2@C80 system was limited to the endohedral arrangement, U2@C80. This restriction is well justified by previous findings by Infante et al.45 that the endohedral bonding of U2 is strongly preferred to the exohedral arrangement in fullerenes C60, C70, and C84. In search for the lowest U2@C80 minimum, local minima were searched by placing the U2 unit (at rU–U = 2.5 Å) in the center of the C80 cage along three different orientations (x, y, or z axis). All seven IPR C80 cages were checked by this procedure. The systems were minimized maintaining the septet electron state44,45 without symmetry constraints. The septet ground state was confirmed by calculating triplet, quintet, and nonet (all geometry optimized). The quality of the unrestricted Kohn–Sham wavefunction was confirmed by negligible spin-contamination, <0.1. The minima were checked by frequency analysis.

The empty C80:7 cage has topological Ih symmetry which is a saddle point due to orbital degeneracy. The empty cage undergoes the Jahn–Teller distortion to a D2 structure.62 For the encapsulation energy calculations we used the C80:7 (Ih) geometry as a starting point and minimized it under D2-symmetry constraints in the singlet ground state.

2.2 QTAIM analysis

The topological properties of critical points within the context of the Quantum Theory of Atoms in Molecules (QTAIM)63 have been employed many times for analyzing the bonding properties in various materials64 including fullerenes, see, e.g.ref. 53 and 65–67.

However, as some of us have shown recently,68 the presence or absence of line critical points (LCP) in a single geometry neither confirm nor invalidate the presence of a chemical bond. In this work, we rely on the profiles of the derivatives of the electron density and a unique quantitative measure of the covalency within the context of QTAIM, delocalization index, δ(A ↔ B) or DI.68

Of the topological profiles, the Laplacian of the electron density, ∇2ρ(r), has been conventionally used to identify the electron density concentration (EDC) between atoms that are believed to be linked to covalency.63 Besides ∇2ρ(r), energy density, H(r), has been proposed to be an efficient tool for distinguishing covalent and polar covalent chemical bonds.69 Energy density at any point in space is defined as H(r) = V(r) + G(r), where V(r) and G(r) are potential and gradient kinetic energy densities. V(r) is always negative at any point in space but G(r) is always positive; a negative H(r) value denotes the dominance of potential energy at a point, which has been interpreted in favour of covalency.

The DI defines the number of electrons that are shared between any pair of atoms,

 
δ(A ↔ B) = −2[〈nAnB〉 − 〈nA〉〈nB〉],(1)
where 〈nAnB〉, 〈nA〉, 〈nB〉 are the localization index or in other word, the minimum electron populations of the atomic pair and single atoms, respectively.

The DI was suggested as a direct measure of electron exchange between atomic basins of two atoms A and B. Recent studies demonstrate that δ(A ↔ B) quantitatively reflects the magnitude of the exchange–correlation energy component for an atomic pair A–B.70,71 The magnitude of DI is close to unity for a typical single homonuclear (sigma) bond, e.g. a carbon–carbon bond in ethane.72 The magnitude of DI for a homonuclear bond reflects the bond order, e.g., it is close to 2, 3, and 4 for double, triple, and quadruple homonuclear bond, respectively. On the contrary, DI of a polar-covalent bond is smaller than the expected value based on the MO picture, which is consistent with chemical intuition for the formation of a polar covalent bond. Nevertheless, it is highly recommended to compare the DI of any system with an external reference to characterize the bond order of a system. Here, we chose U2 as our external reference for assessing the bond order between uranium atoms in the fullerene systems. Scalar-relativistic computations predicted that U2 has a quintuple bond.33 Studying the δ(U ↔ U) of U2 molecule, optimized at the same level of theory as U2@C80, demonstrates that the DI can recover the bond order of this system in a good agreement with previous studies; δ(U ↔ U) = 5.08.

The wavefunction for the analysis of the electron density of the minimum structure was obtained at the BP86/SVP/SDD computational level (cf. above) by Gaussian 09.56 The electron density was analyzed within the context of the QTAIM63 by AIMAll suite of programs.73 For properly treating the uranium atoms in QTAIM analysis, auxiliary basis functions were added to the wavefunction of the molecule.73

3 Results and discussion

3.1 U2C80 is a stable endohedral U2@C80:7 system with a septet ground state. U–U interactions inside the cage are attractive

Endohedral compounds of the M2@C80 formula, so far experimentally observed, are either MIV2C2@C78 systems, such as Ti2C2@C78, or MIII2@C80 systems, such as CeIII2@C80 fullerenes.4,74–76 Because the ionic radii of UIII and CeIII are almost identical, 1.16 Å and 1.15 Å, while those of TiIV and UIV are substantially different, 0.745 Å and 1.03 Å, ref. 77, it can be assumed that an UIII2@C80 system would be formed rather than an UIV2C2@C78 system.

The lowest energy minimum structures for each of the seven possible IPR (isolated pentagon rule)78 C80 cages with enclosed U2, assuming the septet ground state,44–46i.e., six unpaired electrons, are listed in Table 1. The lowest energy minimum derives from the C80:7 cage whereas the optimized minima based on other IPR cages are ca. 10–40 kcal mol−1 less stable. Indeed, the U2C2@C78 isomer was calculated ca. 20 kcal mol−1 higher in energy than the most stable U2@C80 isomer, Table 1.

Table 1 Calculated relative energiesa, ΔE, and U–U distancesb, rU–U, for the IPR-based U2@C80 isomers and U2C2@C78:5 systemc
System ΔE [kcal mol−1] r U–U [Å]
a Relative electronic energies wrt the ground state 7U2@C80:7 calculated at the BP86/SVP/SDD level. b The closest U–Ccage distances are 2.35–2.50 Å. The U–Cendo distances in U2C2@C78 are between 2.20 and 2.30 Å. c Singlet 1U2@C80:7 could not be converged.
7U2@C80:1 42.5 5.117
7U2@C80:2 35.0 5.030
7U2@C80:3 21.2 3.728
7U2@C80:4 22.8 4.198
7U2@C80:5 10.5 3.871
7U2@C80:6 11.8 3.901
3U2@C80:7 22.9 3.965
5U2@C80:7 18.0 3.903
7 U 2 @C 80 :7 0.0 3.894
9U2@C80:7 13.4 3.872
3U2C2@C78:5 20.1 4.256
5U2C2@C78:5 29.8 4.326
7U2C2@C78:5 42.1 4.325


The optimized structure of 7U2@C80:7 is shown in Fig. 1. The molecule has Ci symmetry with the two uranium atoms located nearby a D3 axis of theC80:7 cage. Analogous 3CeIII2@C80:7 has a D3d minimum structure with the cerium atoms and the two closest carbons located on a D3 axis.75 (It is a dynamical system, though.76) Attempted optimization of 7U2@C80:7 within D3d constraints did not converge to a stationary point. At the present level of theory a C2h-symmetric stationary point (one imaginary frequency) could be found about 1 kcal mol−1 above the Ci minimum. With such a small difference, it cannot be excluded that the C2h stationary point becomes the lowest minimum if different computational levels are used.


image file: c5cp04280a-f1.tif
Fig. 1 The 7U2@C80:7 minimum structure calculated at the BP86/SVP/SDD level.

The U–U distance in the 7U2@C80:7 minimum structure of 3.89 Å is rather long as compared to that predicted for U2@C60 (∼2.72–2.74 Å),44–46 or in bare U2 and U22+ (∼2.43 and ∼2.30 Å).33,35,45 It is also longer than twice the empirical single-bond radius of uranium (2 × 1.7 Å = 3.4 Å),79 which suggests the U–U bond order lower than one. However, see below.

The closest U–C bond lengths in 7U2@C80:7 minimum are 2.40, 2.48, and 2.51 Å, comparable to those found in the strongly bound U@C282+, where the rU–C closest contacts are within 2.44–2.51 Å.25 In fact, the cage is significantly stretched along the U–U axis in 7U2@C80:7. The end-to-end distances between the carbon atoms connecting three hexagons on opposite sides of the cage (Fig. 1) vary from 8.16 to 8.33 Å in the empty C80:7, and elongate to 8.68 Å along the U–U axis and to 8.20 Å in the direction perpendicular to the U–U axis in the 7U2@C80:7 minimum structure.

The septet ground state of U2@C80:7 with six unpaired electrons was confirmed by calculating the geometry-optimized nonet, quintet, and triplet, which lay 13, 18, and 23 kcal mol−1 above the ground state septet, Table 1. Note that the previously studied di-uranium fullerenes, U2@C60, U2@C70, U2@C84, and U2@C90, were predicted to be septet in their ground state.44–46 These findings point to a general pattern in the electronic structure of U2@C2n fullerenes. For the future experimental reference, the predicted structure, IR, and Raman spectra of the 7U2@C80:7 lowest minimum structure are given in Table S1 and Fig. S1 and S2 in the ESI.

The encapsulation energy, ΔE, for the 7U2(g) + C80(g) = 7U2@C80:7 reaction was calculated to be −252.7 kcal mol−1 at the BP86/SVP/SDD level. The reaction enthalpy, approximated by the sum of the electronic and zero-point energy was predicted slightly lower, ΔH = −248.6 kcal mol−1. These results are consistent with the previous findings for U2@C60, U2@C70, and U2@C84 where the encapsulation energy was ranging from −160 to −210 kcal mol−1.44,45 For a further comparison, we calculated encapsulation energies for the experimentally known analogous La2@C80 and Ce2@C80 complexes. The calculated ΔE for the M2(g) + C80(g) = M2@C80(g) reaction80,81 is predicted to amount to −255 and −257 kcal mol−1 for M = La, Ce at the BP86/SVP/SDD level.

Interaction between the enclosed uranium atoms in U2@C80 can be actually estimated from a hypothetical isodesmic reaction 2U1@C80 = U2C80 + C80. Thanks to the symmetry of the system, the left side of the reaction corresponds to twice the U–cage interaction and the right side has twice the U–cage plus the U–U interaction in it. The ΔE = −17.7 kcal mol−1 is in favour of products and gives a thermodynamical evidence for endohedral U–U bonding interactions in U2@C80. This evidence is further supported by the bonding analysis given below.

Apparently, the encapsulation energy for U2@C80 of ∼−250 kcal mol−1 is substantially larger than the U–U interaction estimated above, or than the dissociation energy of U2, which was calculated to be −70.1 kcal mol−1 at the BP86/SVP/SDD level or −33.6 kcal mol−1 at the CASPT2 level at corresponding equilibrium distance (rU–U ∼ 2.7 Å).44,45 One may thus expect that the weaker U–U bonding will be strongly affected by the size/type of the cage keeping each uranium atom at a position dictated by the stronger U–C bonding, see Section 3.4.

3.2 MO analysis reveals endohedral U–U bonding orbitals in U2@C80

It is well known that in M2@C80:7 systems, the C80:7 cage formally accepts six electrons from the confined metal atoms.4,65 At the ionic limit, this leads to a C806− cage interacting with two positively charged M3+ ions. However, the electrons are not fully localized on the cage but they are partly shared between the cage and the enclosed cluster.4,65 The bonding situation in 7U2@C80:7 is akin to previously studied M2@C80 compounds. The level of electron transfer is seen on the calculated NPA charge on each uranium atom, qU = +0.82 and on the natural electron configuration (U:7s0.215f3.546d0.98) of enclosed atoms in comparison with the ground state configuration of neutral uranium atoms (U:7s25f36d1). The six unpaired electrons in 7U2@C80:7 are mainly in the U(5f) shell, and are, in fact, strongly localized between the encapsulated uranium atoms as seen from NPA analysis (details not shown) and the frontier orbitals in Fig. 2.
image file: c5cp04280a-f2.tif
Fig. 2 The six highest singly-occupied molecular orbitals (SOMOs) of the 7U2@C80:7 system.

A closer look at the frontier singly-occupied molecular orbitals (SOMO) in Fig. 2 reveals a bonding situation between the enclosed uranium atoms. The SOMO and SOMO−1 orbitals have U–U antibonding character whereas the SOMO−2 through SOMO−5 have U–U bonding character. This situation can be interpreted as two (four bonding minus two antibonding) 1e-2c bonds, in other words a double ferromagnetic bond between the encapsulated uranium atoms. This bond is clearly U(5f)-based. The localization of the unpaired electrons between the uranium atoms is identified also by the calculated spin density in Fig. 3. The localization of the spin density on uranium atoms and its presence on some of the cage carbon atoms may allow for the future experimental identification of the U2@C80 system by ESR49,82 or paramagnetic 13C NMR spectroscopy.83–85


image file: c5cp04280a-f3.tif
Fig. 3 The calculated ground-state spin density for the U2@C80:7 system.

3.3 QTAIM analysis gives evidence for a single U–U bond in U2@C80:7

The bonding between the encapsulated U–U atoms in U2@C80 is further studied by the QTAIM analysis (for details on QTAIM, see Methods section). In this work, rather than on the presence or absence of line critical points we rely on the profiles of the derivatives of the electron density and a unique quantitative measure of the covalency within the context of QTAIM, the delocalization index, δ(A ↔ B).68

Inspecting the Laplacian of the alpha-electron density reveals electron density concentration (EDC) between the uranium atoms, Fig. 4b, a.k.a. U–U bonding interaction. Interestingly, the EDC between the U atoms is not recognizable in the Laplacian of the total electron density, Fig. 4a, since the overall electron density masks the alpha-EDC between the uranium atoms, shown in Fig. 4b. An interesting picture emerges from the Laplacian of the spin density, Fig. 4c, which highlights the regions of spin-density concentration. A profile of f-orbitals and an EDC between two uranium atoms resulting from the f-orbital overlap is rather evident in Fig. 4c. For 3D representations of Fig. 4c, see Fig. S3 (ESI).


image file: c5cp04280a-f4.tif
Fig. 4 Contour maps of (a) Laplacian of electron density, (b) Laplacian of alpha-electron density, (c) Laplacian of spin density, and (d) energy density in the 7U2@C80:7 system; negative functions in all plots, implying electron concentration/sharing are presented by red lines.

In the contour map of energy density, H(r), Fig. 4d, the C–C, U–C and U–U bonds fall in the negative energy density regions, denoting a total stabilization arising from the covalent-type interactions, i.e. electron sharing among the cage carbon atoms, the carbon and the uranium atoms, and between the two uranium atoms.

The delocalization index, DI, for the U–U bond, δ(U ↔ U) was calculated to be 1.01 in 7U2@C80:7. Comparing this value with that for U2 with a quintuple bond calculated at the same level of theory, δ(U ↔ U) = 5.08, suggests that U–U bonding in the U2@C80:7 molecule corresponds to a single U–U bond. This is consistent with the MO picture of two 1e-2c U–U bonds in Fig. 2 above. Notably, calculated δ(M ↔ M) indices for Sc–Sc, Y–Y, Lu–Lu, or La–La interactions in similar dimetallofullerenes were found comparably lower than one,53 within 0.25–0.65, albeit obtained at different levels of theory. The largest value found was δ(Lu ↔ Lu) of 0.65 in Lu2@C82 at rLu–Lu = 3.476 Å. Details can be found in ref. 53.

The energetically higher multiplets of U2@C80:7 (Table 1) give similar DI as that for the septet, δ(U ↔ U) = 1.02 for the triplet and δ(U ↔ U) = 1.37 for the quintet state. Nonet could not be analyzed. The larger DI value for the quintet state is given by the fact that the corresponding electron moves from an U–U alpha antibonding to an U–U beta bonding orbital, while rU–U = 3.90 Å remains similar to that for the septet ground state, where rU–U = 3.89 Å, Table 1.

The magnitude of δ(U ↔ C) for single pair of atoms was found in the range of 0.12 to 0.36 for carbon atoms, which are within the distance of 2.98–2.39 Å from the uranium atoms. These values are within the range of the typical metal–carbon delocalization indices.72

We have thus seen strong theoretical evidence for U–U bonding in the 7U2@C80:7 system, via the attractive U–U potential inside the cage, the presence of a double ferromagnetic bond, or the QTAIM delocalization index for U–U of 1.01. Notably, this U–U bonding is actually observed at relatively large rU–U = 3.89 Å, which is beyond the sum of empirical single-bond radii of uranium of 2 × 1.7 = 3.4 Å.79 The endohedral metal–metal bonding at large M–M separations was recently noted for di-lanthanofullerene anions La2@Cnq, where rLa–La as large as 3.7–5.2 Å gave δ(La ↔ La) = ∼0.3, ref. 53 Hence, we decided to investigate a series of previously studied di-uranofullerenes to estimate how far can the U–U bonding reach and what are the cage-driving capabilities of fullerenes for U–U bonding.

3.4 The U–U bonding along the U2@Cn series reaches beyond U2@C80 and is determined by the cage type

In the following we shortly analyze the U–U bonding interactions and trends in the series of diuranium fullerenes, U2@Cn (n = 60, 70, 80, 84, 90).14,17,44–46 The selection of C70–C90 cages was motivated solely on the basis of previous studies45,46 to illustrate the metal–metal bonding vs. the size of the cage; lower-energy isomers may exist but were not searched for. Optimized structures of U2@Cn are shown in Fig. 5. Selected properties, calculated at the BP86/SVP/SDD level, are summarized in Table 2. We note that only U2@C60 and U2@C80 of the species in Table 2 have been observed experimentally so far.14,17
image file: c5cp04280a-f5.tif
Fig. 5 Optimized structures of U2@C60, U2@C70, U2@C84, and U2@C90. The U–U and closest U–C interactions are shown by dotted lines with the corresponding interatomic distances.
Table 2 The U–U and U–C distances, the encapsulation energies for the 7U2 + Cn7U2@Cn reaction, calculated NPA charge on U atoms, delocalization indices for U–C and U–U bonding, and NPA populations on uranium atoms in the 7U2@Cn (n = 60, 70, 80, 84, 90) EMFs. Calculated at the BP86/SVP/SDD level
System r U–U [Å] r U–C [Å] ΔE [kcal mol−1] q U δ(U ↔ U) [au] δ(U ↔ C) [au] NPA on U
7U2@C60 2.735 2.48–2.49 −200.7 0.07 2.1 4.5 7s0.135f4.186d1.16
7U2@C70 3.923 2.40–2.65 −189.1 0.41 0.7 4.9 7s0.215f3.766d1.03
7U2@C80 3.894 2.40–2.54 −252.7 0.82 1.0 4.1 7s0.215f3.546d0.98
7U2@C84 4.071 2.44–2.67 −152.2 0.78 0.7 4.4 7s0.245f3.626d0.92
7U2@C90 6.358 2.39–2.64 −183.0 0.74 0.1 4.9 7s0.085f3.736d0.88


7U2@C80 has the largest U2 encapsulation energy (−252.7 kcal mol−1) among the studied systems. Generally, the encapsulation energy of diuranium EMFs lies in the range of −150 to −250 kcal mol−1, Table 2.

The U–U distance and the degree of U–U bonding, expressed by the δ(U ↔ U) along the series in Table 2, in fact correlate with the (relative) size of the fullerene cage. No correlation with the encapsulation energies is observed. As noted for U2@C80, the encapsulation energy along the U2@Cn series is substantially larger (150–250 kcal mol−1) than the energy of U–U binding in bare U2 (70.1 kcal mol−1 at the BP86/SVP/SDD level).45 The U–U interaction inside the cage is thus to a large extent dictated by the U–cage bonding. This argument is further supported by rather constant rU–C contact distances along the series, whereas rU–U is changing substantially, as seen in Table 2.

Following the U–U distance and δ(U ↔ U) along the series, we confirm the argument of Infante et al.45 that multiple U–U bonding in U2@C60 is forced by the short U–U distance in the small cage interior. In a large enough cage, like C90, the uranium atoms separate and practically do not interact with each other.46 This is confirmed by negligible δ(U ↔ U) = 0.1 in 7U2@C90. The present results show newly the evidence for the U–U bonding also in cages of intermediate size, U2@C70 through U2@C84, with rU–U ∼ 3.9–4.0 Å, as indicated by the QTAIM delocalization index, δ(U ↔ U) = 0.7–1.0 in Table 2.

The QTAIM analysis shows some general features along the studied series, as is evident from Fig. 6. The EDC, corresponding to the U–U interaction, is absent in the Laplacian of electron densities of U2@C60 and U2@C70 (Fig. 6, panels 1a and 2a) but it appears for that of U2@C84 (Fig. 6, panel 3a). The EDC is also absent in the Laplacian of alpha-ED of U2@C60 (Fig. 6, panel 1b) but appears for the U2@C70 through U2@C84 (Fig. 4b and Fig. 6, panels 2b and 3b). This is due to the masking effect of the electron density of carbon atoms and also the masking of the alpha-ED by beta-ED in the total ED of smaller cages. In all compounds, the Laplacian of the spin density unveils the pattern of f-orbitals involved in accommodating the unpaired 5f-electrons (Fig. 6, panels 1c–4c). The energy-density profiles (Fig. 6, panels 1d–4d) delineate covalently bonded atoms. Notably, the level of U–U interaction can be easily identified qualitatively from the energy density profiles. The strongest effect is seen in U2@C60 and U2@C80, weaker in U2@C70 and U2@C84, and the profile of U2@C90 points to the absence of U–U interaction, see Fig. 6, panels 1d–4d, and Fig. 4d.


image file: c5cp04280a-f6.tif
Fig. 6 Profiles of the Laplacian of (a) total electron density, (b) alpha-electron density, (c) spin density, and (d) the energy density for U2@C60 (1), U2@C70 (2), U2@C84 (3), and U2@C90 (4); red and blue lines depict negative and positive regions of each function, respectively.

The delocalization index serves as a seamless quantitative measure of the order of the U–U bond. With a small deviation for U2@C84 the δ(U ↔ U) decreases as the U–U distance increases in Table 2. In fact, an exponential correlation can be found with a correlation coefficient of r2 = 0.9802 between rU–U inside a cage and δ(U ↔ U), see Fig. 7.


image file: c5cp04280a-f7.tif
Fig. 7 The plot of δ(U ↔ U) versus U–U bond length in U2 (at DI ∼ 5) and along the studied U2@Cn series.

To complete the picture of bonding along the series, we analyzed also the frontier orbitals of the studied compounds, see Fig. S4–S8 in the ESI. The trends in the MO framework along the series are less straightforward than the results of QTAIM analysis and are only discussed briefly. In accord with the work of Wu and Lu44 there are six 1e-2c U–U bonding orbitals (one σ-, three π-, and two δ-orbitals) in U2@C60, see Fig. S4 (ESI). This qualitatively correlates with δ(U ↔ U) = 2.1. In U2@C70 most of the frontier orbitals are actually bonding but they do not overlap efficiently; only two σ-type orbitals show significant U–U overlap, Fig. S5 (ESI), which explain lower δ(U ↔ U) = 0.7 in U2@C70. In U2@C80, we recall the four U–U bonding and two U–U antibonding one-electron orbitals (Fig. 2 and Fig. S6, ESI) and δ(U ↔ U) = 1.0. In U2@C84 there is one σ- and two π-type U–U bonding orbitals (Fig. S7, ESI) but the latter are only weakly overlapping, which explains lower δ(U ↔ U) of 0.7. No U–U bonding orbitals are observed in U2@C90 (Fig. S8, ESI). The definition of the bond order using MO analysis is thus not straightforward. The delocalization index appears as a more genuine and general parameter to be used in the present context.

Finally, the U–U bonding inside a fullerene cage can be regarded as an unwilling bonding. The uranium atoms strongly bind to the cage and acquire a positive charge.53 The calculated NPA charges in Table 2 on uranium atoms vary from 0.1 in U2@C60 and 0.4 in U2@C70 to ∼0.8 in larger fullerenes. To compensate for the U–U charge repulsion, the electron density in U(5f) shells delocalizes between the uranium atoms thus making one-electron-two-center U–U bonds. In a small cage, like U2@C60, covalent multiple U–U bonding with DI(U ↔ U) = 2.1 is forced by the short U–U distance. In larger cages, C70 through C84, the U–U bonding of the order of single bond is still predicted, even at rU–U ∼ 4 Å, with DI(U ↔ U) between 0.7 and 1.0. In U2@C90, strong U–cage interactions and the interior of the cage do not, in principle, prevent U–U bonding but the charge–charge repulsion forces the encapsulated atoms to separate at large distances.

4 Conclusions

In this work, we have theoretically characterized the U2@C80 compound, reported previously in the TOF-MS experiments. We show that it is an endohedral open-shell septet 7U2@C80 system with six unpaired electrons residing mainly in the U(5f) shell. It is derived from the C80:7 fullerene cage. The encapsulated uranium atoms are strongly bound to the cage by polarized covalent bonding, as evidenced by the large encapsulation energy of U2 in C80 and by bonding analysis. MO analysis reveals a double ferromagnetic 1e-2c U–U bond between uranium atoms in the 7U2@C80:7 minimum structure. The U–U bonding is further confirmed by the QTAIM delocalization index, δ(U ↔ U) = 1.01, corresponding to a single bond. This bonding is realized even at a relatively large U–U distance of ∼3.9 Å and arises from U(5f)–U(5f)-based singly occupied MOs. The thermodynamical estimate of the U–U interactions inside 7U2@C80:7 was calculated to be attractive, −17.7 kcal mol−1.

To obtain a more general picture of the endohedral U–U interactions, a series of di-uranium compounds, U2@Cn (n = 60, 70, 80, 84, 90), was analyzed. A U–U bonding of the order of a single bond was also identified in U2@C70 and U2@C84 with rU–U ∼ 4 Å. The character of the U–U bonding and bond order correlates with the U–U distance dictated by the cage, and in this sense can be also tuned by the cage used. The U–U endohedral bonding can be termed as unwilling because it arises from the requirement of the system to decrease the charge–charge repulsion between the encapsulated atoms.

This concept can be extended to other endohedral actinide fullerenes. A preliminary study on the di-thorium fullerenes points to the existence of endohedral Th–Th bonding in C80 and C70 cages. The results will be published elsewhere.

Note added in proof

A new report on experimental preparation of U2@C80 and other di-uranium fullerenes was published recently.86

Acknowledgements

We thank anonymous referees for valuable comments. The project was supported by the Czech Science Foundation, Grant. No. 14-03564S. Institutional support was provided by the Czech Academy of Sciences, project RVO-61388963. C.F.-N. thanks financial support from the SoMoPro II program. The research leading to this work has acquired a financial grant from the People Program (Marie Curie action) of the Seventh Framework Program of EU according to the REA Grant Agreement No. 291782. The research is further co-financed by the South-Moravian Region. Part of the work was carried out at CEITEC – the Central European Institute of Technology with the research infrastructure supported by the project CZ.1.05/1.1.00/02.0068 financed by the European Regional Development Fund. The access to computing and storage facilities owned by parties and projects contributing to the National Grid Infrastructure MetaCentrum, and provided under the program “Projects of Large Infrastructure for Research, Development, and Innovations” (LM2010005), and the CERIT-SC computing and storage facilities, provided under the program Center CERIT Scientific Cloud, part of the Operational Program Research and Development for Innovations (CZ.1.05/3.2.00/08.0144), is acknowledged. We thank Jan Horníček for initial calculations of molecular structures.

References

  1. A. Rodriguez-Fortea, S. Irle and J. M. Poblet, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2011, 1, 350–367 CrossRef CAS PubMed.
  2. H. L. Cong, B. Yu, T. Akasaka and X. Lu, Coord. Chem. Rev., 2013, 257, 2880–2898 CrossRef CAS PubMed.
  3. X. Lu, T. Akasaka and S. Nagase, Acc. Chem. Res., 2013, 46, 1627–1635 CrossRef CAS PubMed.
  4. A. A. Popov, S. Yang and L. Dunsch, Chem. Rev., 2013, 113, 5989–6113 CrossRef CAS PubMed.
  5. T. Wang and C. Wang, Acc. Chem. Res., 2014, 47, 450–458 CrossRef CAS PubMed.
  6. J. Zhang, S. Stevenson and H. C. Dorn, Acc. Chem. Res., 2013, 46, 1548–1557 CrossRef CAS PubMed.
  7. D. M. Rivera-Nazario, J. R. Pinzon, S. Stevenson and L. A. Echegoyen, J. Phys. Org. Chem., 2013, 26, 194–205 CrossRef CAS PubMed.
  8. M. Garcia-Borras, S. Osuna, J. M. Luis, M. Swart and M. Sola, Chem. Soc. Rev., 2014, 43, 5089–5105 RSC.
  9. P. Jin, C. Tang and Z. Chen, Coord. Chem. Rev., 2014, 270, 89–111 CrossRef PubMed.
  10. Y. Zhang and A. A. Popov, Organometallics, 2014, 33, 4537–4549 CrossRef CAS.
  11. P. W. Dunk, M. Mulet-Gas, Y. Nakanishi, N. K. Kaiser, A. Rodriguez-Fortea, H. Shinohara, J. M. Poblet, A. G. Marshall and H. W. Kroto, Nat. Commun., 2014, 5, 8 Search PubMed.
  12. P. Schwerdtfeger, L. N. Wirz and J. Avery, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2015, 5, 96–145 CrossRef CAS PubMed.
  13. J. Zhao, X. Huang, P. Jin and Z. Chen, Coord. Chem. Rev., 2015, 289, 315–340 CrossRef PubMed.
  14. T. Guo, M. D. Diener, Y. Chai, M. J. Alford, R. E. Haufler, S. M. McClure, T. Ohno, J. H. Weaver, G. E. Scuseria and R. E. Smalley, Science, 1992, 257, 1661–1664 CAS.
  15. M. D. Diener, C. A. Smith and D. K. Veirs, Chem. Mater., 1997, 9, 1773–1777 CrossRef CAS.
  16. K. Akiyama, K. Sueki, Y. L. Zhao, H. Haba, K. Tsukada, T. Kodama, K. Kikuchi, T. Ohtsuki, Y. Nagame, H. Nakahara and A. Katada, in Nanonetwork Materials: Fullerenes, Nanotubes and Related Systems, ed. S. Saito, T. Ando, Y. Iwasa, K. Kikuchi, M. Kobayashi and Y. Saito, 2001, vol. 590, pp. 437–440 Search PubMed.
  17. K. Akiyama, Y. L. Zhao, K. Sueki, K. Tsukada, H. Haba, Y. Nagame, T. Kodama, S. Suzuki, T. Ohtsuki, M. Sakaguchi, K. Kikuchi, M. Katada and H. Nakahara, J. Am. Chem. Soc., 2001, 123, 181–182 CrossRef CAS.
  18. K. Akiyama, K. Sueki, K. Tsukada, T. Yaita, Y. Miyake, H. Haba, M. Asai, T. Kodama, K. Kikuchi, T. Ohtsuki, Y. Nagame, M. Katada and H. Nakaharaa, J. Nucl. Radiochem. Sci., 2002, 3, 151–154 CrossRef.
  19. K. Akiyama, K. Sueki, H. Haba, K. Tsukada, M. Asai, T. Yaita, Y. Nagame, K. Kikuchi, M. Katada and H. Nakahara, J. Radioanal. Nucl. Chem., 2003, 255, 155–158 CrossRef CAS.
  20. P. W. Dunk, N. K. Kaiser, M. Mulet-Gas, A. Rodriguez-Fortea, J. M. Poblet, H. Shinohara, C. L. Hendrickson, A. G. Marshall and H. W. Kroto, J. Am. Chem. Soc., 2012, 134, 9380–9389 CrossRef CAS PubMed.
  21. P. W. Fowler, S. J. Austin and J. P. B. Sandall, J. Chem. Soc., Perkin Trans. 2, 1993, 795–797 RSC.
  22. J. Aihara and H. Hosoya, Bull. Chem. Soc. Jpn., 1993, 66, 1955–1958 CrossRef CAS.
  23. K. Jackson, E. Kaxiras and M. R. Pederson, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 17556–17561 CrossRef CAS.
  24. R. M. Pitzer and K. Zhao, J. Phys. Chem., 1996, 100, 4798–4802 CrossRef.
  25. J.-P. Dognon, C. Clavaguera and P. Pyykko, J. Am. Chem. Soc., 2009, 131, 238–243 CrossRef CAS PubMed.
  26. M. V. Ryzhkov, A. L. Ivanovskii and B. Delley, Comput. Theor. Chem., 2012, 985, 46–52 CrossRef CAS PubMed.
  27. D. Manna and T. K. Ghanty, J. Phys. Chem. C, 2012, 116, 25630–25641 CAS.
  28. M. V. Ryzhkov and B. Delley, Comput. Theor. Chem., 2013, 1013, 70–77 CrossRef CAS PubMed.
  29. D. Manna and T. K. Ghanty, J. Phys. Chem. C, 2013, 117, 17859–17869 CAS.
  30. D. Manna, A. Sirohiwal and T. K. Ghanty, J. Phys. Chem. C, 2014, 118, 7211–7221 CAS.
  31. X. Liu, L. Li, B. Liu, D. Wang, Y. Zhao and X. Gao, J. Phys. Chem. A, 2012, 116, 11651–11655 CrossRef CAS PubMed.
  32. F. A. Cotton, D. O. Marler and W. Schwotzer, Inorg. Chem., 1984, 23, 4211–4215 CrossRef CAS.
  33. L. Gagliardi and B. O. Roos, Nature, 2005, 433, 848–851 CrossRef CAS PubMed.
  34. L. N. Gorokhov, A. M. Emelyanov and Y. S. Khodeev, High Temp., 1974, 12, 1156–1158 Search PubMed.
  35. L. Gagliardi, P. Pyykko and B. O. Roos, Phys. Chem. Chem. Phys., 2005, 7, 2415–2417 RSC.
  36. B. O. Roos, P.-A. Malmqvist and L. Gagliardi, J. Am. Chem. Soc., 2006, 128, 17000–17006 CrossRef CAS PubMed.
  37. G. Cavigliasso and N. Kaltsoyannis, Dalton Trans., 2006, 5476–5483 RSC.
  38. G. La Macchia, M. Brynda and L. Gagliardi, Angew. Chem., Int. Ed., 2006, 45, 6210–6213 CrossRef CAS PubMed.
  39. B. O. Roos and L. Gagliardi, Inorg. Chem., 2006, 45, 803–807 CrossRef CAS PubMed.
  40. G. Cavigliasso and N. Kaltsoyannis, Inorg. Chem., 2006, 45, 6828–6839 CrossRef CAS PubMed.
  41. B. O. Roos, A. C. Borin and L. Gagliardi, Angew. Chem., Int. Ed., 2007, 46, 1469–1472 CrossRef CAS PubMed.
  42. G. Cavigliasso and N. Kaltsoyannis, Inorg. Chem., 2007, 46, 3557–3565 CrossRef CAS PubMed.
  43. D. A. Penchoff and B. E. Bursten, Inorg. Chim. Acta, 2015, 424, 267–273 CrossRef CAS PubMed.
  44. X. Wu and X. Lu, J. Am. Chem. Soc., 2007, 129, 2171–2177 CrossRef CAS PubMed.
  45. I. Infante, L. Gagliardi and G. E. Scuseria, J. Am. Chem. Soc., 2008, 130, 7459–7465 CrossRef CAS PubMed.
  46. X. Dai, Y. Meng, M. Xin, F. Wang, D. Fei, M. Jin, Z. Wang and R. Zhang, Procedia Chem., 2012, 7, 528–533 CrossRef CAS PubMed.
  47. X. Dai, J. Han, Y. Gao and Z. G. Wang, ChemPhysChem, 2014, 15, 3871–3876 CrossRef CAS PubMed.
  48. J. Han, X. Dai, Y. Gao, Y. Meng and Z. Wang, Phys. Chem. Chem. Phys., 2014, 16, 22784–22790 RSC.
  49. T. Zuo, L. Xu, C. M. Beavers, M. M. Olmstead, W. Fu, D. Crawford, A. L. Balch and H. C. Dorn, J. Am. Chem. Soc., 2008, 130, 12992–12997 CrossRef CAS PubMed.
  50. H. Umemoto, K. Ohashi, T. Inoue, N. Fukui, T. Sugai and H. Shinohara, Chem. Commun., 2010, 46, 5653–5655 RSC.
  51. T. Yang, X. Zhao and E. Osawa, Chem. – Eur. J., 2011, 17, 10230–10234 CrossRef CAS PubMed.
  52. J. Y. Hao, F. Y. Li, H. J. Li, X. Y. Chen, Y. Y. Zhang, Z. F. Chen and C. Hao, RSC Adv., 2015, 5, 34383–34389 RSC.
  53. A. A. Popov, S. M. Avdoshenko, A. Martin Pendas and L. Dunsch, Chem. Commun., 2012, 48, 8031–8050 RSC.
  54. C. Schäfer, C. Huber and R. Ahlrichs, J. Chem. Phys., 1994, 100, 5829 CrossRef PubMed.
  55. X. Cao and M. Dolg, J. Mol. Struct., 2004, 673, 203–209 CrossRef CAS PubMed.
  56. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö Farkas, J. B. Foresman and J. V. Ortiz, J. Cioslowski, D. J. Fox, Gaussian 09, Revision D.01, Wallingford CT, 2009 Search PubMed.
  57. R. Ahlrichs, M. Bär, M. Häser, H. Horn and C. Kölmel, Chem. Phys. Lett., 1989, 162, 165–169 CrossRef CAS.
  58. A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899–926 CrossRef CAS.
  59. A. E. Reed, R. B. Weinstock and F. Weinhold, J. Chem. Phys., 1985, 83, 735–746 CrossRef CAS PubMed.
  60. T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
  61. W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics Modell., 1996, 14, 33–38 CrossRef CAS.
  62. M.-L. Sun, Z. Slanina, S.-L. Lee and F. Uhlík, Chem. Phys. Lett., 1995, 246, 66–72 CrossRef CAS.
  63. R. W. F. Bader, Atoms in Molecules: A Quantum Theory, Oxford University Press, Oxford, 1990 Search PubMed.
  64. P. L. A. Popelier, Atoms in Molecules an Introduction, Pearson, London, 2000 Search PubMed.
  65. A. A. Popov and L. Dunsch, Chem. – Eur. J., 2009, 15, 9707–9729 CrossRef CAS PubMed.
  66. Z. Badri, C. Foroutan-Nejad and P. Rashidi-Ranjbar, Comput. Theor. Chem., 2013, 1009, 103–107 CrossRef CAS PubMed.
  67. Y. Kolokoltsev, D. P. Sanders and V. A. Basiuk, J. Comput. Theor. Nanosci., 2015, 12, 674–681 CrossRef CAS PubMed.
  68. C. Foroutan-Nejad, S. Shahbazian and R. Marek, Chem. – Eur. J., 2014, 20, 10140 CrossRef CAS PubMed.
  69. D. Cremer and E. Kraka, Angew. Chem., Int. Ed., 1984, 23, 627–628 CrossRef PubMed.
  70. M. Rafat and P. L. A. Popelier, in The quantum theory of atoms in molecules: from solid state to DNA and drug design., ed. C. F. Matta and R. J. Boyed, Willey-VCH, Weinheim, 2007, p. 121 Search PubMed.
  71. M. García-Revilla, E. Francisco, P. L. A. Popelier and A. Martín-Pendás, ChemPhysChem, 2013, 14, 1211–1218 CrossRef PubMed.
  72. F. Cortés-Guzmán and R. F. W. Bader, Coord. Chem. Rev., 2005, 249, 633–662 CrossRef PubMed.
  73. T. A. Keith, AIMAll, version 13.05.06, http://aim.tkgristmill.com Search PubMed.
  74. T. Yumura, Y. Sato, K. Suenaga and S. Iijima, J. Phys. Chem. B, 2005, 109, 20251–20255 CrossRef CAS PubMed.
  75. K. Muthukumar and J. A. Larsson, J. Mater. Chem., 2008, 18, 3347–3351 RSC.
  76. L. Feng, M. Suzuki, N. Mizorogi, X. Lu, M. Yamada, T. Akasaka and S. Nagase, Chem. – Eur. J., 2013, 19, 988–993 CrossRef CAS PubMed.
  77. J. E. Huheey, E. A. Keiter and L. L. Keiter, Inorganic Chemistry: Principles of Structure and Reactivity, Fourth Edition, HarperCollins College Publishers, New York, 1993 Search PubMed.
  78. P. W. Fowler and D. E. Manolopoulos, An Atlas of Fullerenes, Oxford University Press, Oxford, U.K., 1995 Search PubMed.
  79. P. Pyykkö, J. Phys. Chem. A, 2015, 119, 2326–2337 CrossRef PubMed.
  80. X. Y. Cao, W. J. Liu and M. Dolg, Sci. China, Ser. B: Chem., 2002, 45, 91–96 CrossRef CAS.
  81. X. Y. Cao and M. Dolg, Theor. Chem. Acc., 2002, 108, 143–149 CrossRef CAS.
  82. T. Kato, J. Mol. Struct., 2007, 838, 84–88 CrossRef CAS PubMed.
  83. M. Yamada, T. Wakahara, Y. Lian, T. Tsuchiya, T. Akasaka, M. Waelchli, N. Mizorogi, S. Nagase and K. M. Kadish, J. Am. Chem. Soc., 2006, 128, 1400–1401 CrossRef CAS PubMed.
  84. S. Taubert, M. Straka, T. O. Pennanen, D. Sundholm and J. Vaara, Phys. Chem. Chem. Phys., 2008, 10, 7158–7168 RSC.
  85. Y. Zhang, D. Krylov, M. Rosenkranz, S. Schiemenz and A. A. Popov, Chem. Sci., 2015, 6, 2328–2341 RSC.
  86. L. Echegoyen and D. Buck, in 227th ECS Meeting, Chicago, IL, 2015, abstract 817 Search PubMed.

Footnote

Electronic supplementary information (ESI) available: Table S1 with xyz coordinates of U2@C80 lowest minimum and Fig. S1 and S2 with IR and Raman spectra of U2@C80 lowest minimum. Fig. S3 with Laplacian of spin density for U2@C80 and Fig. S4–S8 with frontier molecular orbitals for 7U2@Cn (n = 60, 70, 80, 84, 90) compounds. See DOI: 10.1039/c5cp04280a

This journal is © the Owner Societies 2015