Next Article in Journal
Composition-Dependent Phonon and Thermodynamic Characteristics of C-Based XxY1−xC (X, Y ≡ Si, Ge, Sn) Alloys
Previous Article in Journal
The Effect of Ge Doping on α-Ag2S’s Thermoelectric and Mechanical Properties
Previous Article in Special Issue
Endometallofullerenes in the Gas Phase: Progress and Prospect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Structural and Vibrational Properties of Solid Endohedral Metallofullerene Li@C60

1
Division of Materials Physics, Ruđer Bošković Institute, 10000 Zagreb, Croatia
2
Center for High-Pressure Science & Technology Advanced Research, Beijing 100094, China
3
Department of Mechanical Engineering and Materials Science and Engineering, Cyprus University of Technology, Limassol 3036, Cyprus
4
Department of Physical Science and Engineering, Nagoya Institute of Technology, Nagoya 466-8555, Japan
5
Department of Physics, Graduate School of Science, Osaka Metropolitan University, Osaka 599-8531, Japan
6
Department of Materials Science, Graduate School of Engineering, Osaka Metropolitan University, Osaka 599-8531, Japan
7
Division of Applied Chemistry, Graduate School of Engineering, Osaka University, Osaka 565-0871, Japan
8
RIKEN SPring-8 Center, 1-1-1 Kouto, Sayo-gun, Sayo-cho 679-5148, Japan
9
Japan Synchrotron Radiation Research Institute (JASRI), 1-1-1 Kouto, Sayo-gun, Sayo-cho 679-5198, Japan
10
Physics Department, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
11
Faculty of Engineering, Kyoto University of Advanced Science, Kameoka 621-8555, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Present address: Nano Carbon Device Research Center, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba 305-8565, Japan.
Inorganics 2024, 12(4), 99; https://doi.org/10.3390/inorganics12040099
Submission received: 3 March 2024 / Revised: 28 March 2024 / Accepted: 28 March 2024 / Published: 29 March 2024
(This article belongs to the Special Issue Research on Metallofullerenes)

Abstract

:
The endohedral lithium fulleride, Li+@C60•−, is a potential precursor for new families of molecular superconducting and electronic materials beyond those accessible to date from C60 itself. Solid Li@C60 comprises (Li@C60)2 dimers, isostructural and isoelectronic with the (C59N)2 units found in solid azafullerene. Here, we investigate the structural and vibrational properties of Li@C60 samples synthesized by electrolytic reduction routes. The resulting materials are of high quality, with crystallinity far superior to that of their antecedents isolated by chemical reduction. They permit facile, unambiguous identification of both the reduced state of the fulleride units and the interball C-C bonds responsible for dimerization. However, severe orientational disorder conceals any crystal symmetry lowering due to the presence of dimers. Diffraction reveals the adoption of a hexagonal crystal structure (space group P63/mmc) at both low temperatures and high pressures, typically associated with close-packing of spherical monomer units. Such a situation is reminiscent of the structural behavior of the high-pressure Phase I of solid dihydrogen, H2.

1. Introduction

Following the serendipitous discovery of the fullerene family of molecular carbon nearly 40 years ago, research on the chemical, physical, and materials properties of its members and those of their derivatives has proceeded unabated. In brief, there are four distinct ways of manipulating the molecular and crystalline forms of the fullerenes. Firstly, direct chemical functionalization of the π-electron molecular framework leads to a rich organic chemistry, affording numerous derivatives of both C60 and its higher fullerene antecedents that have applications in diverse fields from molecular electronics to solar cells [1].
Secondly, reductive intercalation of metal ions into the voids of the crystalline structures of fullerenes—most prominently, that of C60—leads to sensitively modifying the electronic structure and has resulted in a plethora of fulleride salts with MxC60 stoichiometry and x ranging from a low value of 1 (e.g., in RbC60) to a high value of 12 (e.g., in Li12C60). In particular, exohedrally modified fulleride materials with A3C60 (A = alkali metal) stoichiometry have provided examples of high-Tc molecular superconductors with transition temperatures as high as 38 K and with the zero-resistance performance surviving to record high magnetic fields (exceeding 90 T). Of significance is also the unconventional, highly correlated nature of superconductivity, which is proximate to an antiferromagnetic Mott insulating state out of which it emerges upon lattice contraction via the application of chemical and/or physical pressure [2,3,4,5].
Thirdly, the electronic structure can be tuned by on-ball substitution of heteroatoms such as B or N. The most prominent derivative has been the azafullerene radical, C59N, which dimerizes to afford the non-magnetic (C59N)2 moiety [6]. Further manipulation of the properties of such heterofullerenes can be achieved by exohedral metal intercalation, resulting in salts such as K6C59N, in which electron doping is achieved synergistically both by on- and in-between-the-balls donors [7].
Finally, a large family of fullerene derivatives can be accessed by endohedral encapsulation of atomic or molecular units to afford A@Cn moieties. For a long while, this field of fullerene research was dominated by metal encapsulation of higher fullerenes [8,9], with bulk availability of endohedral C60 units remaining elusive. This situation has changed dramatically following two significant discoveries: (a) the development of the “molecular surgery” synthetic technique [10], which has led to numerous endohedral C60 fullerenes incorporating noble gas atoms (e.g., He, Ne, Ar, Kr) or small molecules (e.g., H2, HF, CO, H2O, CH4) [11,12] and (b) the successful synthesis of fullerene salts, (Li+@C60)X (X = PF6, SbCl6), which comprise Li+@C60 endohedral molecular ions and X spacer units, which serve a charge-balancing role [13,14,15]. Remarkably, both chemical and electrochemical reduction of the Li+@C60 cation lead to the formation of the neutral radical anion, Li+@C60•− [16,17], which, in analogy with the on-ball-substituted azafullerene C59N radical, dimerizes to afford (Li@C60)2, isostructural and isoelectronic with the (C59N)2 molecule [18,19].
Structural and electronic characterization of the (Li@C60)2 solid obtained by chemical reduction confirmed its dimeric structure in the solid state [17]. However, the as-obtained material lacked good crystallinity, resulting in moderate quality diffraction and spectroscopic data. Attempts to sublime the material led to the escape of the metal ion from the endofullerene space, precluding any improvement in crystallinity, in contrast to its azafullerene antecedent [19]. Here, we report on the structural and vibrational properties of the (Li@C60)2 solid obtained by electrochemical reduction—this is a highly crystalline material, which permits detailed study of the structural and electronic properties. While vibrational spectroscopy readily identifies the vibrational signatures of interball C-C bond formation consistent with dimerization, the material is shown to adopt a highly symmetric hexagonal-close-packed crystal structure—typically associated with monomer sphere packing—as a result of disorder both at ambient pressure down to 3.4 K and at ambient temperature up to 13.1 GPa.

2. Results and Discussion

2.1. Vibrational Properties

The Raman spectrum of (Li@C60)2 obtained by electrochemical reduction of the (Li+@C60)(PF6) precursor, along with those of monomeric C60, (Li@C60)(PF6), and dimeric (C59N)2 solids, is illustrated in Figure 1—the corresponding molecular units are generated by VESTA [20]. The frequencies of the observed Raman peaks are summarized in Table S1. In the case of the monomeric C60 and (Li@C60)(PF6) fullerene materials, the 785 nm laser line was used for excitation to prevent photoinduced polymerization upon illumination with visible or ultraviolet light [21]. On the other hand, owing to the stability of the dimeric (Li@C60)2 and (C59N)2 fullerenes against photoinduced polymerization, the 633 nm line of a He-Ne laser was used for their excitation to take advantage of the improved throughput of the Raman spectrometer in the corresponding spectral region. The Raman spectra of C60 solid and (Li@C60)(PF6) are nearly identical to each other with small frequency shifts (≤3 cm−1) between the corresponding peaks and are compatible with the monomeric nature of the C60 units. Both spectra are dominated by the ten Raman-active intramolecular C60 modes, attributed to the two totally symmetric Ag(1) and Ag(2) modes and the eight fivefold-degenerate Hg(1)-Hg(8) of the C60 molecule in icosahedral Ih symmetry [22]. The weak crystal field in solid C60 causes only the splitting of the radial Hg(1) mode into two components (at 266 and 272 cm−1), while the appearance of a weak peak at 1458 cm−1 can be attributed to the presence of traces of polymerized C60 (Ag(2) mode of polymeric chains [23]).
The Raman spectrum of dimeric (C59N)2 solid is far richer in structure compared to the monomeric systems, as the cage distortion and the concomitant lowering of the molecular symmetry upon dimerization cause the appearance of additional Raman peaks. These originate from the splitting of the degenerate Hg modes and the activation and splitting of inactive vibrational modes of the C60 molecule in Ih symmetry [24,25]. The plethora of Raman split components appearing in the spectrum of (C59N)2 is in fair agreement with those earlier reported in the literature and assigned to intramolecular C60 modes in Ih symmetry [24]. The same assignment of the observed Raman peaks in the present work was also adopted in Table S1. As inferred from Figure 1 and Table S1, the similarity of the Raman spectrum profile and the frequencies of the constituent Raman peaks of (Li@C60)2 obtained by electrochemical reduction with those of dimeric (C59N)2 is evident, establishing its dimeric nature unambiguously. Moreover, comparison of the overall Raman spectrum profile of the present (Li@C60)2 sample with that of the neutral Li@C60 obtained by chemical reduction of the Li+@C60 cation by decamethylferrocene (Figure S1) clearly illustrates the stronger intensities and narrower lineshapes of the split Raman components in (Li@C60)2 obtained by electrochemical reduction—this is in addition to the similarities regarding the appearance of the Raman peaks associated with the dimerization of the C60 cages. The present results are compatible with the superior crystallinity of the present material, in accordance with its powder XRD data (vide infra), contrary to the situation encountered in (Li@C60)2 obtained by chemical reduction, where a small fraction of monomers is also present [17].
The frequency of the Ag(2) pentagonal pinch mode of the C60 molecule deserves further attention. In particular, the Ag(2) frequency in C60-based materials is extremely sensitive to both doping and intermolecular covalent bonding [23,24,25,26,27,28]. In monomeric alkali fullerides, it downshifts quasi-linearly by 6–7 cm−1 per electron transferred from the metal atoms to the C60 units [26,27,28], due to the elongation of the intramolecular bonds induced by the charge transfer process and the resulting softening of the force constants [29,30]. Moreover, in neutral polymeric C60-based fullerene materials—including dimerized systems—the Ag(2) mode also softens due to the intermolecular covalent bond formation and the concomitant decrease in the mean intramolecular bond strength. The observed softening depends strongly on the number of the sp3-like coordinated carbon atoms per molecular cage [31]. Accordingly, and as evident in Table S1 (numbers in bold), there is a softening of the Ag(2) peak in neutral (C59N)2 compared to that in C60 solid by ~6 cm−1, which is attributed to the formation of the intermolecular covalent bond due to dimerization [25]. In the case of (Li@C60)2, the softening of this peak is even larger, ~13 and ~10 cm−1 relative to monomeric C60 and (Li@C60)(PF6), respectively. This can be understood in terms of the combined effect of covalent intermolecular bonding and uptake of one electron by the C60 cage in (Li@C60)2—here, electrons are more delocalized over the radical carbanion cage than in the case of (C59N)2, where the electron pairs are localized on the nitrogen atoms [17].
The dimeric nature of (Li@C60)2 obtained by electrochemical reduction is further supported by its IR transmission spectrum illustrated in Figure 2 in comparison with that of monomeric C60 solid. More specifically, the IR spectrum of C60 solid is dominated by the four IR active modes of the C60 molecule in Ih symmetry {F1u(1)-F1u(4)}, appearing at 525, 576, 1182, and 1428 cm−1, in excellent agreement with the literature [32]. The additional weaker spectral features are due to forbidden fundamentals, activated by the distortion of the molecular Ih symmetry in the C60 crystal. As in the case of the corresponding Raman spectrum, in the IR spectrum of (Li@C60)2, several split components appear due to intercage bonding and symmetry lowering, similar to the case of (C59N)2 [33]. Despite similarities of the IR spectrum with that of monomeric C59HN, the appearance of characteristic features in the frequency region 800–850 cm−1 (marked accordingly in Figure 2) provides a fingerprint of the dimerization of the molecular cages [33].

2.2. Structural Properties at Ambient Pressure

The synchrotron X-ray powder diffraction patterns of the (Li@C60)2 sample recorded at 300 and 3.4 K at ambient pressure are shown in Figure 3. An immediate observation is that they appear strikingly similar to those reported for as-prepared solid (C59N)2, in terms of both the position and the relative intensity of the diffraction peaks present [18,34]. As-prepared (C59N)2 crystallizes in a hexagonal lattice (space group P63/mmc) with lattice constants at ambient temperature, a = 9.97 Å and c = 16.18 Å [18]. In a remarkable analogy, the patterns of (Li@C60)2 prepared by electrochemical reduction are also consistent with the material adopting an isostructural hexagonal closed-packed (hcp) arrangement of the fullerene units with a somewhat smaller (c/a) anisotropy (space group P63/mmc, at 300 K: a = 10.04374(10) Å and c = 16.13451(19) Å; at 3.4 K: a = 9.9560(3) Å and c = 16.0317(5) Å). Excellent refinements of the diffraction profiles are achieved by the LeBail pattern–decomposition technique using the GSAS suite (at 300 K: Rwp = 1.88%, Rexp = 1.48%; at 3.4 K: Rwp = 2.23%, Rexp = 1.43%) (Figure 3).
On the other hand, the measured profiles of the present (Li@C60)2 sample contrast sharply with those of both (Li@C60)2 prepared by chemical reduction route [17] and sublimed (C59N)2 [19]. The former displays a powder pattern, which can be indexed on a pseudo-cubic unit cell with a lattice size of ~14.15 Å (interfullerene separation ~10.01 Å) [17], whereas the latter provides clear evidence of a dimerized molecular structure and a lowering of crystal symmetry to monoclinic [19].
The apparent hcp crystal structure (Figure 3a inset) of (Li@C60)2 prepared by electrochemical reduction is an archetypal packing arrangement of monomer units. Therefore, it is counterintuitive and contrasts sharply with the dimerization of the molecular units that has been unambiguously established by vibrational spectroscopy (vide supra). In the P63/mmc space group, the Li@C60 balls reside in the 2d positions, (⅔, ⅓, ¼) and (⅓, ⅔, ¾). As a result, their center-to-center distance at ambient temperature is 9.94 Å, implicit of the absence of interball bond formation. Due to the translational symmetry of the hexagonal unit cell, there are six nearest neighbors at this distance of 9.94 Å, together with six additional ones in the ab pane at 10.04 Å. These are close to the C60—C60 distances in monomeric C60 (~10 Å) and are longer than those (~9.34 Å) in typical dimerized C60 structures like that of RbC60 [35,36,37].
The inconsistency with the vibrational results may be resolved if we introduce orientational disorder of the dimer units residing on the hcp sites. To this effect, we recall a much-celebrated analogous situation—namely, that of the Phase I of solid dihydrogen, H2 [38]. At room temperature, this phase is stable between 5 and 190 GPa and comprises freely rotating dihydrogen molecules, forming an hcp crystal structure (space group P63/mmc). We therefore attempted Rietveld refinements of the diffraction data by modeling the pronounced orientational disorder of the fullerene moieties in terms of spherical shell scatterers placed at the 2d sites of 6 ¯ m2 symmetry in the hcp lattice. In the course of the refinements, we found it necessary to also introduce scattering density at the 2a sites of 3 ¯ m. symmetry, presumably reflecting the presence of residual CH2Cl2 solvent molecules intercalated during the electrochemical synthetic protocol. Rietveld refinements now proceeded smoothly, resulting in a stoichiometry of (Li@C60)2⋅(CH2Cl2)0.81(2) for the material. We note that an analogous situation was encountered in the Rietveld refinements of the diffraction data of as-prepared hcp-structured azafullerene solid, whose stoichiometry was established as (C59N)2⋅(CS2)1.0 [34]. The refinement results are shown in Figure 4 and Table S2.
Figure 5a,b shows the temperature dependence of the unit–cell volume and the lattice parameters a and c, as extracted from the LeBail analyses of the diffraction data. The smooth evolution in lattice size confirms the absence of any structural phase transition down to 3.4 K. The only apparent effect is a decrease in the lattice anisotropy, as measured by the (c/a) ratio, which monotonically increases with decreasing temperature, approaching a value just larger than 1.61 at 3.4 K (Figure 5c). This is still smaller than the ideal value of 1.633 in perfect hexagonal close-packed architectures, but the decreased anisotropy on cooling may reflect a partial freezing out of the orientational disorder of the fullerene units.

2.3. Structural Properties at High Pressure

Synchrotron X-ray powder diffraction data of (Li@C60)2 were collected at ambient temperature and at selected pressures between 0.5 and 13.1 GPa. As the pressure increases, the diffraction profiles show gradual Bragg peak shifting towards higher 2θ angles consistent with unit cell contraction; this is accompanied by some peak broadening due to increasing inhomogeneities upon pressurization (Figure 6). No evidence of the occurrence of any structural phase transition is found in the pressure range of the present experiments. All diffraction patterns index with the same hexagonal unit cell (space group P63/mmc) found at all temperatures at ambient pressure. Representative LeBail refinements at base and highest pressures are included in Figure 6 (at 0.5 GPa: a = 9.89824(18) Å and c = 15.9333(3) Å; Rwp = 0.35%, Rexp = 4.60%; at 13.1 GPa: a = 9.2495(5) Å and c = 15.0437(11) Å; Rwp = 0.29%, Rexp = 4.94%). The evolution of the (Li@C60)2 unit cell dimensions upon pressurization was investigated by LeBail refinements of the diffraction profiles (Figure 7a,b). Some spurious weak peaks, which were present in the diffraction datasets when the sample was introduced into the diamond anvil cell (DAC)—but were absent for the same sample batch outside the pressure cell—were excluded from the refinements. Figure 7c displays the change in the lattice anisotropy of the hcp structure, as monitored by the evolution of the (c/a) ratio. Upon pressurization, c/a increases monotonically at the beginning but eventually saturates above ~10 GPa towards a value of ~1.626, close to the ideal value of 1.633 expected for perfect hcp stacking of solid spheres. Comparable behavior was also observed before for the isostructural dimeric (C59N)2 solid, in which the (c/a) ratio approached ~1.633 at ~6.5 GPa [18]. Such behavior implies that, at these elevated pressures, the intradimer and interdimer C-C distances in the compressed material have become comparable. Notably, this occurs in the same pressure range that the interdimer center-to-center separations in (Li@C60)2 have decreased to values on the order of ~9.2–9.3 Å. Such close interball contacts have provided the signature of single C-C bond formation in fullerene-derived polymers [35,36,37] before.
The pressure dependence of the unit cell volume of (Li@C60)2 was next investigated using the semi-empirical second-order Murnaghan equation-of-state (EoS):
P = K 0 K 0 [ V 0 V K 0 1 ]
where K0 is the atmospheric pressure isothermal bulk modulus, K0′ is its pressure derivative (≡ dK0/dP)P=0, and V0 is the unit–cell volume at zero pressure (approximated by the ambient pressure datum at 1 atm; here, V0 = 1409.58 Å3). The EoS fit to the V(P) data (see Equation (S1) in the Supplementary Materials) of (Li@C60)2 up to a pressure of 13.1 GPa results in values of K0 = 10.8(3) GPa (leading to an initial volume compressibility, κ = −dlnV/dP = 0.093(3) GPa−1) and of K0′ = 11.4(2) (Figure 7a). The measured volume compressibility of (Li@C60)2 is higher than those of (C59N)2 (κ = 0.046(2) GPa−1) and of the monomeric C60 (κ = 0.055 GPa−1) and (Li+@C60)(PF6) (κ = 0.0571(4) GPa−1) [15,18,39] solids, implying more flexible nearest-neighbor bonds and somewhat less tight crystal packing. Modified EoS equations, in which K0 and its pressure derivative, K0′, were replaced by Kx and Kx′ (x = a or c) parameters, were used to fit the corresponding lattice axes data to extract initial axial compressibilities, βx ( = 1/(3Kx) GPa−1, x = a or c) (Figure 7b). As anticipated by the earlier discussion on the pressure dependence of the (c/a) ratio, we find that the compressibility along the c-axis (βc = 0.0086(2) GPa−1) is smaller than that along the a-axis (βa = 0.0112(3) GPa−1). Such a pressure response was also observed for (C59N)2 and was then interpreted as a consequence of the intradimer bonds being oriented along the (⅓, ⅔, ½) direction and therefore inclined at a larger angle to the basal plane than to the c-axis [18].

3. Materials and Methods

The (Li+@C60)(PF6) sample used as a precursor in the present work was purchased from Idea International Corporation. Its electrochemical reduction to afford (Li@C60)2 was performed according to the experimental protocol described in the literature before [16].
Raman spectra of (Li@C60)2 were recorded in backscattering geometry using a LabRam HR (HORIBA, Kyoto, Japan) micro-Raman spectrometer, equipped with a Peltier cooled CCD detector. Laser lines at 633 and 785 nm were alternatively used for excitation, focused on the samples by means of a 100× objective at a power lower than 0.3 mW to avoid any laser-heating-induced effects.
Infrared (IR) transmission spectra of the samples under argon atmosphere were collected by means of a JASCO (Tokyo, Japan) FT-IR 6200 spectrometer.
Synchrotron X-ray powder diffraction measurements of (Li@C60)2 at various temperatures between 3.4 and 300 K at ambient pressure were performed with the high-resolution powder diffractometer on beamline BL44B2 at SPring-8, Hyogo, Japan using a helium-flow cryostat [40]. The powder sample was sealed in a 0.5 mm diameter thin-wall glass capillary under He gas. The wavelength was 0.800312(1) Å. LeBail and Rietveld analyses were performed with the GSAS and Fullprof suites of refinement programs [41,42,43].
Synchrotron X-ray powder diffraction measurements of (Li@C60)2 up to 13.1 GPa at ambient temperature were conducted at the BL10XU beamline, SPring-8, Hyogo, Japan. The powder X-ray diffraction data (monochromatized beam with λ = 0.41277(5) Å, beam size = 40 μm) were collected on an image plate detector (Rigaku (Tokyo, Japan) R-AXIS IV++, 300 × 300 mm2, and 0.10 mm pixel size) while oscillating the sample [44]. Masking of the Bragg reflections of the DAC and integration of the 2D diffraction images to afford one-dimensional diffraction profiles were processed with the IPanalyzer software [45]. LeBail analysis of the diffraction profiles was performed with the GSAS suite of refinement programs [41,42]. A DAC [46,47], equipped with single-crystal diamonds of 500 μm culet diameter and with an 80 μm-thick stainless-steel gasket, was used for pressure generation. Powder samples were introduced into 250 μm diameter holes made at the center of the gasket, together with a ruby sphere and helium gas (solidification pressure ∼12.1 GPa at room temperature, pressure uncertainty less than 0.5% at the highest pressure of the present experiment [48]) as pressure marker and pressure medium, respectively. The applied pressure was measured by the ruby–fluorescence method [49]. All measured pressure values are quoted to one decimal point.

4. Conclusions

In conclusion, we have presented a detailed structural and vibrational characterization of the dimerized endohedral metallofullerene, (Li@C60)2, in the solid state. The material used in the present work was obtained by electrochemical reduction of the trifluoromethanesulfonyl imide precursor salt, (Li+@C60)(TFSI) [16]. Its highly crystalline nature permits facile identification of spectroscopic features related to interfullerene C-C bonding. Remarkably, the crystal structure of the material remains hexagonal close-packed (space group P63/mmc) both at low temperatures and ambient pressure and at high pressures and ambient temperature, concealing the effects of dimerization and the anticipated lowering in symmetry. High orientational disorder (together with the presence of some residual dichloromethane solvent) can account for the apparent high symmetry. A comparable orientational disorder scenario has been used to rationalize the hcp Phase I of solid dihydrogen (H2) at high pressures.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics12040099/s1, Figure S1: Comparison of Raman spectra of (Li@C60)2 obtained by chemical and electrochemical reduction; Table S1: Assignment of observed Raman peaks and their frequencies for C60, (Li@C60)(PF6), (C59N)2, and (Li@C60)2; Table S2: Refined parameters of the Rietveld analysis of the synchrotron X-ray powder diffraction data of (Li@C60)2; Equation (S1): V(P) form of the Murnaghan equation-of-state.

Author Contributions

Synthesis, M.V., T.N. and K.K. (Ken Kokubo); synchrotron X-ray diffraction—data collection and analysis, M.V., T.N., M.M., Y.T., K.K. (Kenichi Kato) and K.P.; high-pressure synchrotron X-ray diffraction—data collection and analysis, N.Y., K.M., S.K.-I., H.K. and K.P.; vibrational spectroscopy, J.A.; conceptualization and overall conclusions, J.A., Y.K. and K.P.; project supervision, K.P.; writing—review and editing, K.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by Grants-in-Aid for Scientific Research (JSPS KAKENHI grant numbers JP18K05254, JP21H01907, JP22K18693, and JP23KJ1843) and by JST SPRING, grant number JPMJSP2139.

Data Availability Statement

All data supporting the reported results in this paper can be obtained from the authors.

Acknowledgments

We thank SPring-8 for access to synchrotron XRPD measurements (proposal no. 2017A4131, 2017A1363, 2017B1276, 2017B1358, 2018A1398, and 2022A1699).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hirsch, A.; Brettreich, M. Fullerenes: Chemistry and Reactions; Wiley-VCH, 5 Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2005. [Google Scholar]
  2. Takabayashi, Y.; Prassides, K. Unconventional high-Tc superconductivity in fullerides. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20150320. [Google Scholar] [CrossRef] [PubMed]
  3. Zadik, R.H.; Takabayashi, Y.; Klupp, G.; Colman, R.H.; Ganin, A.Y.; Potocnik, A.; Jeglic, P.; Arcon, D.; Matus, P.; Kamaras, K.; et al. Optimized unconventional superconductivity in a molecular Jahn-Teller metal. Sci. Adv. 2015, 1, e1500059. [Google Scholar] [CrossRef]
  4. Nomura, Y.; Sakai, S.; Capone, M.; Arita, R. Unified understanding of superconductivity and Mott transition in alkali-doped fullerides from first principles. Sci. Adv. 2015, 1, e1500568. [Google Scholar] [CrossRef] [PubMed]
  5. Kasahara, Y.; Takeuchi, Y.; Zadik, R.H.; Takabayashi, Y.; Colman, R.H.; McDonald, R.D.; Rosseinsky, M.J.; Prassides, K.; Iwasa, Y. Upper critical field reaches 90 tesla near the Mott transition in fulleride superconductors. Nat. Commun. 2017, 8, 14467. [Google Scholar] [CrossRef] [PubMed]
  6. Hummelen, J.C.; Knight, B.; Pavlovich, J.; Gonzalez, R.; Wudl, F. Isolation of the heterofullerene C59N as its dimer (C59N)2. Science 1995, 269, 1554–1556. [Google Scholar] [CrossRef] [PubMed]
  7. Prassides, K.; Keshavarz-K, M.; Hummelen, J.C.; Andreoni, W.; Giannozzi, P.; Beer, E.; Bellavia, C.; Cristofolini, L.; Gonzalez, R.; Lappas, A.; et al. Isolation, structure, and electronic calculations of the heterofullerene salt K6C59N. Science 1996, 271, 1833–1835. [Google Scholar] [CrossRef]
  8. Shinohara, H.; Tagmatarchis, N. Endohedral Metallofullerenes: Fullereness with Metal Inside; Wiley: Chichester, UK, 2015. [Google Scholar]
  9. Akasaka, T.; Nagase, S. Endofullerenes: A New Family of Carbon Clusters; Springer: Dordrecht, The Netherlands, 2002. [Google Scholar]
  10. Komatsu, K.; Murata, M.; Murata, Y. Encapsulation of molecular hydrogen in fullerene C60 by organic synthesis. Science 2005, 307, 238–240. [Google Scholar] [CrossRef] [PubMed]
  11. Bloodworth, S.; Whitby, R.J. Synthesis of endohedral fullerenes by molecular surgery. Commun. Chem. 2022, 5, 121. [Google Scholar] [CrossRef] [PubMed]
  12. Krachmalnicoff, A.; Bounds, R.; Mamone, S.; Alom, S.; Concistrè, M.; Meier, B.; Kouřil, K.; Light, M.E.; Johnson, M.R.; Rols, S.; et al. The dipolar endofullerene HF@C60. Nat. Chem. 2016, 8, 953–957. [Google Scholar] [CrossRef]
  13. 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]
  14. 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. 2012, 124, 3433–3437. [Google Scholar] [CrossRef]
  15. Colman, R.H.; Okur, H.E.; Garbarino, G.; Ohishi, Y.; Aoyagi, S.; Shinohara, H.; Prassides, K. Pressure effects on the crystal structure of the cubic metallofullerene salt [Li@C60][PF6] to 12 GPa. Mater. Today Commun. 2022, 31, 103275. [Google Scholar] [CrossRef]
  16. Ueno, H.; Aoyagi, S.; Yamazaki, Y.; Ohkubo, K.; Ikuma, N.; Okada, H.; Kato, T.; Matsuo, Y.; Fukuzumi, S.; Kokubo, K. Electrochemical Reduction of Cationic Li+@C60 to Neutral Li+@C60˙: Isolation and Characterisation of Endohedral [60] Fulleride. Chem. Sci. 2016, 7, 5770–5774. [Google Scholar] [CrossRef] [PubMed]
  17. Okada, H.; Ueno, H.; Takabayashi, Y.; Nakagawa, T.; Vrankic, M.; Arvanitidis, J.; Kusamoto, T.; Prassides, K.; Matsuo, Y. Chemical reduction of Li+@C60 by decamethylferrocene to produce neutral Li+@C60•−. Carbon 2019, 153, 467–471. [Google Scholar] [CrossRef]
  18. Brown, C.M.; Beer, E.; Bellavia, C.; Cristofolini, L.; González, R.; Hanfland, M.; Häusermann, D.; Keshavarz-K, M.; Kordatos, K.; Prassides, K.; et al. Effects of Pressure on the Azafullerene (C59N)2 Molecular Solid to 22 GPa. J. Am. Chem. Soc. 1996, 118, 8715–8716. [Google Scholar] [CrossRef]
  19. Brown, C.M.; Cristofolini, L.; Kordatos, K.; Prassides, K.; Bellavia, C.; González, R.; Keshavarz-K, M.; Wudl, F.; Cheetham, A.K.; Zhang, J.P.; et al. On the Crystal Structure of Azafullerene (C59N)2. Chem. Mater. 1996, 8, 2548–2550. [Google Scholar] [CrossRef]
  20. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  21. Rao, A.M.; Zhou, P.; Wang, K.-A.; Hager, G.T.; Holden, J.M.; Wang, Y.; Lee, W.-T.; Bi, X.-X.; Ecklund, P.C.; Cornett, D.S.; et al. Photoinduced polymerization of solid C60 films. Science 1993, 259, 955–957. [Google Scholar] [CrossRef]
  22. Bethune, D.S.; Meijer, G.; Tang, W.C.; Rosen, H.J.; Golden, W.G.; Seki, H.; Brown, C.A.; de Vries, M.S. Vibrational Raman and infrared spectra of chromatographically separated C60 and C70 fullerene clusters. Chem. Phys. Lett. 1991, 179, 181–186. [Google Scholar] [CrossRef]
  23. Davydov, V.A.; Kashevarova, L.S.; Rakhmanina, A.V.; Senyavin, V.M.; Ceolin, R.; Szwarc, H.; Allouchi, H.; Agafonov, V. Spectroscopic study of pressure-polymerized phases of C60. Phys. Rev. B 2000, 61, 11936–11945. [Google Scholar] [CrossRef]
  24. Kuzmany, H.; Plank, W.; Winter, J.; Dubay, O.; Tagmatarchis, N.; Prassides, K. Raman spectrum and stability of (C59N)2. Phys. Rev. B 1999, 60, 1005–1012. [Google Scholar] [CrossRef]
  25. Plank, W.; Pichler, T.; Kuzmany, H.; Dubay, O.; Tagmatarchis, N.; Prassides, K. Resonance Raman excitation and electronic structure of the single bonded dimers (C60)2 and (C59N)2. Eur. Phys. J. B 2000, 17, 33–42. [Google Scholar] [CrossRef]
  26. Winter, J.; Kuzmany, H. Potassium doped fullerene KxC60 with x = 0, 1, 2, 3, 4 and 6. Solid State Commun. 1992, 84, 935–938. [Google Scholar] [CrossRef]
  27. Kuzmany, H.; Matus, M.; Burger, B.; Winter, J. Raman scattering in C60 fullerenes and fullerides. Adv. Mater. 1994, 6, 731–745. [Google Scholar] [CrossRef]
  28. Kosaka, M.; Tanigaki, K.; Prassides, K.; Margadonna, S.; Lappas, A.; Brown, C.M.; Fitch, A.N. Superconductivity in LixCsC60 fullerides. Phys. Rev. B 1999, 59, R6628–R6630. [Google Scholar] [CrossRef]
  29. Jishi, R.A.; Dresselhaus, M.S. Mode softening and mode stiffening in C60 doped with alkali metals. Phys. Rev. B 1992, 45, 6914–6918. [Google Scholar] [CrossRef] [PubMed]
  30. Zhou, P.; Wang, K.A.; Wang, Y.; Eklund, P.C.; Dresselhaus, M.S.; Dresselhaus, G.; Jishi, R.A. Raman scattering in C60 and alkali-metal-saturated C60. Phys. Rev. B 1992, 46, 2595–2605. [Google Scholar] [CrossRef] [PubMed]
  31. Meletov, K.P.; Velkos, G.; Arvanitidis, J.; Christofilos, D.; Kourouklis, G.A. Raman study of the photopolymer formation in the {Pt(dbdtc)2}·C60 fullerene complex and the decomposition kinetics of the photo-oligomers. Chem. Phys. Lett. 2017, 68, 124–129. [Google Scholar] [CrossRef]
  32. Krause, M.; Dunsch, L.; Seifert, G.; Fowler, P.W.; Gromov, A.; Kratschmer, W.; Gutierrez, R.; Porezag, D.; Frauenheim, T.J. Vibrational signatures of fullerene oxides. J. Chem. Soc. Faraday Trans. 1998, 94, 2287–2294. [Google Scholar] [CrossRef]
  33. Krause, M.; Baes-Fischlmair, S.; Pfeiffer, R.; Plank, W.; Pichler, T.; Kuzmany, H.; Tagmatarchis, N.; Prassides, K. Thermal Stability and high temperature graphitization of bisazafullerene (C59N)2 as studied by IR and Raman spectroscopy. J. Phys. Chem. B 2001, 105, 11964–11969. [Google Scholar] [CrossRef]
  34. Prassides, K.; Wudl, F.; Andreoni, W. Solid Azafullerenes and Azafullerides. Fullerene Sci. Technol. 1997, 5, 801–812. [Google Scholar] [CrossRef]
  35. Oszlányi, G.; Bortel, G.; Faigel, G.; Gránásy, L.; Bendele, G.M.; Stephens, P.W.; Forró, L. Single C-C Bond in (C60)22−. Phys. Rev. B 1996, 54, 11849–11852. [Google Scholar] [CrossRef]
  36. Prassides, K.; Vavekis, K.; Kordatos, K.; Tanigaki, K.; Bendele, G.M.; Stephens, P.W. Loss of cubic symmetry in low-temperature Na2RbC60. J. Am. Chem. Soc. 1997, 119, 834–835. [Google Scholar] [CrossRef]
  37. Bendele, G.M.; Stephens, P.W.; Prassides, K.; Vavekis, K.; Kordatos, K.; Tanigaki, K. Effect of charge state on polymeric bonding geometry: The ground state of Na2RbC60. Phys. Rev. Lett. 1998, 80, 736–739. [Google Scholar] [CrossRef]
  38. Ji, C.; Li, B.; Liu, W.; Smith, J.S.; Majumdar, A.; Luo, W.; Ahuja, R.; Shu, J.; Wang, J.; Sinogeikin, S.; et al. Ultrahigh-pressure isostructural electronic transitions in hydrogen. Nature 2019, 573, 558–562. [Google Scholar] [CrossRef]
  39. Duclos, S.J.; Brister, K.; Haddon, R.C.; Kortan, A.R.; Thiel, F.A. Effects of Pressure and Stress on C60 Fullerite to 20 GPa. Nature 1991, 351, 380–382. [Google Scholar] [CrossRef]
  40. Kato, K.; Tanaka, H. Visualizing charge densities and electrostatic potentials in materials by synchrotron X-ray powder diffraction. Adv. Phys. X 2016, 1, 55–80. [Google Scholar] [CrossRef]
  41. Larson, A.C.; von Dreele, R.B. General Structure Analysis System (GSAS). In Los Alamos National Laboratory Report LAUR; The Regents of the University of California: Oakland, CA, USA, 2000; pp. 86–748. [Google Scholar]
  42. Toby, B.H. EXPGUI, a Graphical User Interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  43. Rodríguez-Carvajal, J. Recent Advances in Magnetic Structure Determination by Neutron Powder Diffraction. Physica B Condens. Matter 1993, 192, 55–69. [Google Scholar] [CrossRef]
  44. Hirao, N.; Kawaguchi, S.I.; Hirose, K.; Shimizu, K.; Ohtani, E.; Ohishi, Y. New developments in high-pressure X-ray diffraction beamline for diamond anvil cell at SPring-8. Matter Radiat. Extrem. 2020, 5, 018403. [Google Scholar] [CrossRef]
  45. Seto, Y.; Nishio-Hamane, D.; Nagai, T.; Sata, N. Development of a software suite on X-ray diffraction experiments. Rev. High Press. Sci. Technol. 2010, 20, 269–276. [Google Scholar] [CrossRef]
  46. Sakai, T.; Yagi, T.; Ohfuji, H.; Irifune, T.; Ohishi, Y.; Hirao, N.; Suzuki, Y.; Kuroda, Y.; Asakawa, T.; Kanemura, T. High-pressure generation using double stage micro-paired diamond anvils shaped by focused ion beam. Rev. Sci. Instrum. 2015, 86, 033905. [Google Scholar] [CrossRef] [PubMed]
  47. Sakai, T.; Yagi, T.; Irifune, T.; Kadobayashi, H.; Hirao, N.; Kunimoto, T.; Ohfuji, H.; Kawaguchi-Imada, S.; Ohishi, Y.; Tateno, S.; et al. High pressure generation using double-stage diamond anvil technique: Problems and equations of state of rhenium. High Press. Res. 2018, 38, 107–119. [Google Scholar] [CrossRef]
  48. Zha, C.-S.; Mao, H.-K.; Hemley, R.J. Elasticity of MgO and a Primary Pressure Scale to 55 GPa. Proc. Natl. Acad. Sci. USA 2000, 97, 13494. [Google Scholar] [CrossRef]
  49. Klotz, S.; Chervin, J.C.; Munsch, P.; Le Marchand, G. Hydrostatic limits of 11 pressure transmitting media. J. Phys. D Appl. Phys. 2009, 42, 075413. [Google Scholar] [CrossRef]
Figure 1. The Raman spectrum of the (Li@C60)2 sample prepared by electrochemical reduction (λexc = 633 nm) is displayed together with those of C60exc = 785 nm), monomeric (Li@C60)(PF6) (λexc = 785 nm), and dimeric (C59N)2exc = 633 nm) solids. The low-frequency spectral region in the dimeric molecular solids has been suitably enhanced to improve visibility. The mode assignment refers to the irreducible representations of the C60 molecule in Ih symmetry. Each spectrum is accompanied by the structure of the corresponding molecular unit drawn by VESTA [20].
Figure 1. The Raman spectrum of the (Li@C60)2 sample prepared by electrochemical reduction (λexc = 633 nm) is displayed together with those of C60exc = 785 nm), monomeric (Li@C60)(PF6) (λexc = 785 nm), and dimeric (C59N)2exc = 633 nm) solids. The low-frequency spectral region in the dimeric molecular solids has been suitably enhanced to improve visibility. The mode assignment refers to the irreducible representations of the C60 molecule in Ih symmetry. Each spectrum is accompanied by the structure of the corresponding molecular unit drawn by VESTA [20].
Inorganics 12 00099 g001
Figure 2. IR transmission spectrum of the (Li@C60)2 sample prepared by electrochemical reduction and compared to that of solid C60. Characteristic “dimer peaks” in the frequency region 800–850 cm−1 are indicated. The mode assignment refers to the irreducible representations of the C60 molecule in Ih symmetry.
Figure 2. IR transmission spectrum of the (Li@C60)2 sample prepared by electrochemical reduction and compared to that of solid C60. Characteristic “dimer peaks” in the frequency region 800–850 cm−1 are indicated. The mode assignment refers to the irreducible representations of the C60 molecule in Ih symmetry.
Inorganics 12 00099 g002
Figure 3. LeBail refinements (using the GSAS suite of programs) of the synchrotron X-ray (λ = 0.80031 Å) powder diffraction profiles of (Li@C60)2 at (a) 300 K and (b) 3.4 K. The red open circles represent the measured intensity, and the blue solid line shows the calculated intensity. The lower green solid line shows the difference profile, and the tick marks show the reflection positions in space group P63/mmc. The inset in (a) shows a schematic diagram of the apparent hexagonal close-packing motif of solid spheres, consistent with the space group symmetry.
Figure 3. LeBail refinements (using the GSAS suite of programs) of the synchrotron X-ray (λ = 0.80031 Å) powder diffraction profiles of (Li@C60)2 at (a) 300 K and (b) 3.4 K. The red open circles represent the measured intensity, and the blue solid line shows the calculated intensity. The lower green solid line shows the difference profile, and the tick marks show the reflection positions in space group P63/mmc. The inset in (a) shows a schematic diagram of the apparent hexagonal close-packing motif of solid spheres, consistent with the space group symmetry.
Inorganics 12 00099 g003
Figure 4. Rietveld analysis result (space group P63/mmc) using the Fullprof suite of programs—final observed (red circles) and calculated (blue solid line) synchrotron X-ray (λ = 0.80031 Å) powder diffraction profiles for the (Li@C60)2 sample at 300 K. The lower solid line shows the difference profile, and the tick marks show the reflection positions. Refined parameters and agreement indices are provided in Supplementary Information, Table S2.
Figure 4. Rietveld analysis result (space group P63/mmc) using the Fullprof suite of programs—final observed (red circles) and calculated (blue solid line) synchrotron X-ray (λ = 0.80031 Å) powder diffraction profiles for the (Li@C60)2 sample at 300 K. The lower solid line shows the difference profile, and the tick marks show the reflection positions. Refined parameters and agreement indices are provided in Supplementary Information, Table S2.
Inorganics 12 00099 g004
Figure 5. Temperature dependence of (a) the unit cell volume, (b) the hexagonal lattice constants, a and c, and (c) the (c/a) ratio in (Li@C60)2 at ambient pressure. The lines through the points are guides-to-the-eye.
Figure 5. Temperature dependence of (a) the unit cell volume, (b) the hexagonal lattice constants, a and c, and (c) the (c/a) ratio in (Li@C60)2 at ambient pressure. The lines through the points are guides-to-the-eye.
Inorganics 12 00099 g005
Figure 6. Representative high-pressure synchrotron X-ray powder diffraction profiles after background subtraction (λ = 0.4128 Å) of (Li@C60)2 (red solid lines and open circles) at elevated pressures. The corresponding diffraction profile at 1 atm (capillary) from Figure 5 is also displayed for comparison after conversion to the same wavelength (black solid line). The LeBail refinement results for base (0.5 GPa) and highest (13.1 GPa) pressure are also included (blue solid lines: calculated, red circles: observed). Tick marks label the reflections of the hexagonal lattice (space group P63/mmc), and the green solid lines are the difference profiles between observed and calculated profiles. Some spurious weak peaks present in the high-pressure data with the sample inside the diamond anvil cell (DAC) are excluded from the refinements.
Figure 6. Representative high-pressure synchrotron X-ray powder diffraction profiles after background subtraction (λ = 0.4128 Å) of (Li@C60)2 (red solid lines and open circles) at elevated pressures. The corresponding diffraction profile at 1 atm (capillary) from Figure 5 is also displayed for comparison after conversion to the same wavelength (black solid line). The LeBail refinement results for base (0.5 GPa) and highest (13.1 GPa) pressure are also included (blue solid lines: calculated, red circles: observed). Tick marks label the reflections of the hexagonal lattice (space group P63/mmc), and the green solid lines are the difference profiles between observed and calculated profiles. Some spurious weak peaks present in the high-pressure data with the sample inside the diamond anvil cell (DAC) are excluded from the refinements.
Inorganics 12 00099 g006
Figure 7. Pressure dependence of (a) the unit cell volume, (b) the hexagonal lattice constants, a and c, and (c) the (c/a) ratio in (Li@C60)2 at ambient temperature. The solid lines in (a,b) are the results of least-squares fits to the semi-empirical 2nd-order Murnaghan equation-of-states (EoS) and its axial compressibility variants, respectively. The line through the points in (c) is a guide-to-the-eye.
Figure 7. Pressure dependence of (a) the unit cell volume, (b) the hexagonal lattice constants, a and c, and (c) the (c/a) ratio in (Li@C60)2 at ambient temperature. The solid lines in (a,b) are the results of least-squares fits to the semi-empirical 2nd-order Murnaghan equation-of-states (EoS) and its axial compressibility variants, respectively. The line through the points in (c) is a guide-to-the-eye.
Inorganics 12 00099 g007
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

Vrankić, M.; Nakagawa, T.; Menelaou, M.; Takabayashi, Y.; Yoshikane, N.; Matsui, K.; Kokubo, K.; Kato, K.; Kawaguchi-Imada, S.; Kadobayashi, H.; et al. On the Structural and Vibrational Properties of Solid Endohedral Metallofullerene Li@C60. Inorganics 2024, 12, 99. https://doi.org/10.3390/inorganics12040099

AMA Style

Vrankić M, Nakagawa T, Menelaou M, Takabayashi Y, Yoshikane N, Matsui K, Kokubo K, Kato K, Kawaguchi-Imada S, Kadobayashi H, et al. On the Structural and Vibrational Properties of Solid Endohedral Metallofullerene Li@C60. Inorganics. 2024; 12(4):99. https://doi.org/10.3390/inorganics12040099

Chicago/Turabian Style

Vrankić, Martina, Takeshi Nakagawa, Melita Menelaou, Yasuhiro Takabayashi, Naoya Yoshikane, Keisuke Matsui, Ken Kokubo, Kenichi Kato, Saori Kawaguchi-Imada, Hirokazu Kadobayashi, and et al. 2024. "On the Structural and Vibrational Properties of Solid Endohedral Metallofullerene Li@C60" Inorganics 12, no. 4: 99. https://doi.org/10.3390/inorganics12040099

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

Article Metrics

Back to TopTop