Next Article in Journal
Parameters Synthesis of Na-Magadiite Materials for Water Treatment and Removal of Basic Blue-41: Properties and Single-Batch Design Adsorber
Next Article in Special Issue
Process Parameter Optimisation for Endohedral Metallofullerene Synthesis via the Arc-Discharge Method
Previous Article in Journal
Theoretical Investigation of Inorganic Particulate Matter: The Case of Water Adsorption on a NaCl Particle Model Studied Using Grand Canonical Monte Carlo Simulations
Previous Article in Special Issue
Reactivity of Open-Shell Metallofullerene Anions: Synthesis, Crystal Structures, and Electrochemical Properties of Benzylated Gd@C2v-C82
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of a Novel Non-Isolated-Pentagon-Rule Isomer of Th@C76:Th@C1(17418)-C76

1
College of Chemistry, Chemical Engineering and Materials Science, and State Key Laboratory of Radiation Medicine and Protection, Soochow University, Suzhou 215123, China
2
CAS Key Laboratory of Materials for Energy Conversion, Anhui Laboratory of Advanced Photon Science and Technology, Department of Materials Science and Engineering, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Inorganics 2023, 11(11), 422; https://doi.org/10.3390/inorganics11110422
Submission received: 19 September 2023 / Revised: 16 October 2023 / Accepted: 24 October 2023 / Published: 25 October 2023
(This article belongs to the Special Issue Research on Metallofullerenes)

Abstract

:
A novel Non-Isolated-Pentagon-Rule (non-IPR) isomer of thorium-based endohedral mono-metallofullerenes (mono-EMFs), Th@C1(17418)-C76, was successfully synthesized and characterized using MALDI-TOF mass spectroscopy, single-crystal X-ray diffraction, UV-vis-NIR spectroscopy, and Raman spectroscopy. The molecular structure of this non-IPR isomer was determined unambiguously as Th@C1(17418)-C76 using a single-crystal X-ray diffraction analysis. The crystallographic results further revealed that the optimal Th site resided at the intersection of two adjacent pentagons, similar to that of U@C1(17418)-C76. Additionally, the UV-vis-NIR spectra of Th@C1(17418)-C76 exhibited distinct differences compared to the previously reported U@C1(17418)-C76, highlighting the distinctive electronic structure of actinium-based endohedral metallofullerenes (EMFs). The Raman spectrum of Th@C1(17418)-C76 exhibited similarities to that previously reported for thorium-based EMFs, indicating the analogous strong metal–cage interactions of thorium-based EMFs.

Graphical Abstract

1. Introduction

Endohedral metallofullerenes (EMFs) are a specialized type of carbon nanomaterial featured by the encapsulation of metal atoms or metal clusters within the carbon cage of the fullerenes [1,2,3]. In 1991, Smalley et al. achieved the isolation of the first EMF, La@C82, which initiated the exploration of the structure and properties of EMFs [4]. Subsequently, numerous types of mono-EMFs and endohedral clusterfullerenes have been synthesized and reported by researchers [5,6,7,8]. Depending on the differences between endohedral species, EMFs can be further classified as mono-metallofullerenes, di-metallofullerenes, tri-metallofullerenes, and clusterfullerenes [9,10,11,12,13]. The electron transfer and host–guest interactions between the endohedral metal units and fullerene carbon cages contribute to the intriguing electronic structures and physicochemical properties of these EMFs, thereby presenting immense potential in various cutting-edge fields, including biomedicine, organic photovoltaic devices, and single-molecule devices [14,15,16,17].
Mono-EMFs represent the most extensively studied type of EMF, with the majority of reported mono-EMFs being based on lanthanide metals. These metals are embedded within the carbon cage of the fullerene in the form of M@C2n (M = Ce [18], Pr [19], Nd [20], Sm [21], Eu [22], Gd [23], Tb [24], Dy [25], Ho [26], Er [27], Yb [28], and Lu [29]). In recent years, research on actinium-based mono-EMFs has also been reported, primarily focused on uranium (U) and thorium (Th). Among them, the characterization of Th@C3v(8)-C82 [30] marked the beginning of the investigation into actinium-based mono-EMFs. In monometallic thorium-based EMFs, thorium usually exhibits an oxidation state of a +4 oxidation state. For example, Th@Td(19151)-C76 [31], Th@D5h(6)-C80 [32], Th@C2v(9)-C82, Th@C2(5)-C82 [33], and Th@C1(11)-C86 [34] have been discovered in previous research. All of them demonstrate a four-electron charge transfer from the metal atom to the fullerene cage. Conversely, uranium’s oxidation state displays a remarkable flexibility. In U@C2v(9)-C82, the endohedral uranium transfers three electrons to the C82 carbon cage, while in U@C2(5)-C82, the endohedral uranium transfers four electrons to the C82 fullerene cage [35]. Because of the contributions from f-orbital electrons, the unique characteristics and variability of oxidation states in actinide metals lead to more complex electron configurations in actinide-based EMFs compared to lanthanide-based EMFs. Actinide-based EMFs also exhibit stronger metal–cage interactions than those of lanthanide-based EMFs, and actinide metal ions have a more significant influence on the properties of actinide-based EMFs than previously observed for lanthanide-based EMFs. For example, previously reported mono-EMFs based on uranium, known as U@C1(28324)-C80, and mono-EMFs based on thorium, labeled as Th@C1(28324) C80, possess the same fullerene carbon cage, and both of them undergo a four-electron charge transfer process. They demonstrate distinct electronic properties, as evidenced by significant differences in their UV-vis-NIR spectra [36].
Based on extensive research on actinium-based mono-EMFs, it is clear that the remarkable charge transfer and strong engagement between actinide metal ions and fullerene carbon cages are essential for stabilizing previously unstable fullerene cage structures. Thus, in addition to the experimentally verified isomers, there are numerous isomers that have only been predicted and researched theoretically in actinium-based mono-EMFs. Consequently, further experimental investigations are of paramount importance to exploring actinium-based mono-EMFs, as they will reveal the distinct electronic properties and carbon cage structures that differentiate them from traditional lanthanide-based mono-EMFs.
Recently, two independent computational studies proposed two possible isomeric structures of Th@C76, one is referred to as Th@Td(19151)-C76, while the other is referred to as Th@C1(17418)-C76. [37,38]. Nevertheless, only the existence of Th@Td(19151)-C76 [31] has been experimentally confirmed. Further experimental investigation of Th@C1(17418)-C76 would contribute to a deeper understanding of the unique carbon cage structures, electronic properties, and host–guest interactions exhibited by actinide-based EMFs.
In this study, we present the synthesis and characterization of a novel non-IPR isomer of thorium-based mono-EMFs, Th@C1(17418)-C76. Th@C1(17418)-C76 was successfully synthesized by using a modified Krätschmer–Huffman direct-current DC arc-discharge method. It was thoroughly characterized through various techniques, including MALDI-TOF mass spectroscopy, single-crystal X-ray diffraction, UV-vis-NIR spectroscopy, and Raman spectroscopy. The crystallographic results revealed that the optimal Th site resided at the intersection of two adjacent pentagons, and the presence of a single Th ion played a crucial role in stabilizing the non-IPR C76 fullerene cage.

2. Results and Discussion

2.1. Synthesis and Isolation of Th@C76

Thorium-based EMFs were produced using a modified Krätschmer–Huffman direct-current DC arc-discharge method. Hollow graphite rods, packed with ThO2 and graphite powder (molar ratio of Th: C = 1:24), were vaporized in an arcing reactor chamber under a 200 Torr helium atmosphere. The resulting soot was then collected and extracted with CS2 for 12 h. Multistage high-performance liquid chromatography (HPLC) separation processes were employed to isolate and purify the Th@C76 (Figure S1). The purity of the Th@C76 was confirmed by the observation of single peaks with HPLC and positive-ion-mode matrix-assisted laser desorption-ionization time-of-flight (MALDI-TOF) mass spectrometry, as exhibited in Figure 1. The mass spectra of the isomer of Th@C76 exhibited a peak at 1143.975 m/z, and the experimental isotopic distribution of the sample agrees well with the empirical projection.

2.2. Molecular Structure of Th@C1(17418)-C76

The molecular configuration of Th@C76 was unequivocally ascertained through single-crystal X-ray diffraction analyses. Th@C76 and NiII-OEP (OEP = 2, 3, 7, 8, 12, 13, 17, 18-octaethylporphyrin dianion) were co-crystallized by slowly diffusing the benzene solution of NiII-OEP into the carbon disulfide solution of the purified compounds. The structure of Th@C76 was resolved and refined in the C2/c (No. 15) space group. The experimental results clearly showed that this new isomer of Th@C76 showed a different structure from the reported Th@Td(19151)-C76 [31]. It shares the same carbon cage structure with U@C1(17418)-C76, which possesses a low-symmetry C1(17418)-C76 cage and one adjacent pentagon pair [36]. Interestingly, crystallographic investigations of the C1(17418)-C76 carbon cage have primarily been limited to actinide metallofullerenes, and no previous studies have reported the observation of or noticed the same cage in lanthanide-based EMFs. This indicates that a stronger interaction between U or Th and fullerene cages may facilitate the stabilization of these unconventional fullerene cage isomers. On the other hand, Th@C1(17418)-C76 was also predicted to be the most advantageous product in experiments at elevated temperatures due to its dominant molar fraction [38]. This result verifies the theoretical predications.
Figure 2 shows the molecular structure of Th@C1(17418)-C76 and its relationship to the co-crystallized NiII(OEP) molecule. The Th@C1(17418)-C76 moiety is close to the adjacent NiII(OEP), with a short Ni-to-cage carbon distance (Ni1-C12) of 2.875 Å, demonstrating a characteristic interplay between the carbon cage and NiII(OEP). There are five crystallographic sites for the Th atoms with occupancies of 0.50 for Th1, 0.30 for Th2, 0.14 for Th3, 0.03 for Th4, and 0.025 for Th5 (see Figure S2). Th1 (occupancy 0.50) is situated at the intersection of two adjacent pentagons with the shortest Th−C distance in the range of 2.359−2.449 Å (Table S2), as shown in Figure 3. The shortest Th−C distances of the Th@C1(17418)-C76 are similar to those in Th@Td(19151)-C76, demonstrating similar interactions between the Th atom and the carbon cage [31]. In comparison to two examples of lanthanide-based non-IPR monometallic C76 (M@C2v(19138)-C76, M = Eu [39], Sm [40]), although the major metal sites are all located at the intersection of two adjacent pentagons, there are large differences in the shortest Th–cage distances. This discrepancy suggests that the interaction between the metal and cage in Th@C76 could potentially possess a greater strength when compared to its lanthanide counterparts.

2.3. Spectroscopic Characterizations

UV-vis-NIR absorption is a useful technique that can help us to determine the electronic structures of EMFs. Normally, EMFs sharing the same cage and charge transfer exhibit almost identical absorption spectra [1]. The UV-vis-NIR absorption spectrum was utilized in this study to characterize the pure sample dissolved in a CS2 solution, as depicted in Figure 4. This allowed for further analysis. Th@C1(17418)-C76 exhibits five characteristic peaks at 490, 602, 755, 930, and 1078 nm. The absorption onset occurs at approximately 1230 nm, indicating an optical band of 1.01 V. The absorption spectrum of U@C1(17418)-C76 [36], which shares the same cage isomer and charge transfer, though showing some similarity to that of the Th@C1(17418)-C76, only exhibits one absorption at 593 nm (Figure 4a). This result shows a notable influence of the actinide metal on the electronic structures of M@C1(17418)-C76(M = U, Th). A similar phenomenon can also be observed in M@C1(28324)-C80 [36] (M = U, Th), revealing that the endohedral metal itself also plays a significant role in the electronic structure of the EMFs. On the other hand, the spectrum of Th@Td(19151)-C76 [31] shows different characteristic peaks compared to those of the Th@C1(17418)-C76 (Figure 4b). This agrees with the crystallographic results that the Th@C76 identified in this work is a novel isomer. To delve deeper into the electronic structure of Th@C1(17418)-C76, we conducted electrochemical tests on the compound. The cyclic voltammogram of Th@C1(17418)-C76 presents two oxidation peaks and three reduction peaks. However, the cyclic voltammogram of U@ C1(17418)-C76 only presents one oxidation peak and three reduction peaks [36]. In addition, compared to the first oxidation potential of U@C1(17418)-C76 (0.14 V), the first oxidation potential of Th@C1(17418)-C76 (0.37 V) shows a significant positive shift. This result shows the notable influence of the actinide metal on the electronic structures of M@C1(17418)-C76(M = U, Th).
Th@C1(17418)-C76 was also characterized using low-energy Raman spectroscopy, as shown in Figure 5. Sharp peaks at 149, 218, and 479 cm−1 are observed for the Raman spectrum of Th@C1(17418)-C76. Similar frequencies also can be observed in Th@D5h(6)-C80 [32] (150 and 219 cm−1), Th@C2v(9)-C82 (151 and 216 cm−1), and Th@C2(5)-C82 [33] (152 and 215 cm−1), indicating that thorium-based mono-EMFs have similar metal–cage interactions. Though the UV-vis-NIR absorption spectra of Th@C1(17418)-C76 differ from those of U@C1(17418)-C76, the vibrational frequencies on the Raman spectrum of U@C1(17418)-C76 are at 153, 219, and 359 cm−1 [36], showing notable similarities to those of the Th@C1(17418)-C76. This phenomenon can also be observed in Th@C1(28134)-C80 and U@C1(28134)-C80 [36], which also share the same carbon cage. This may indicate that the cage isomers play a major role on the metal–cage interaction of the corresponding EMFs.

3. Materials and Methods

Materials. The chemicals employed in this research were Th2O3 (99.9%, School of Radiation Medicine and Protection of Soochow University., Suzhou, China), CS2 (99%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), toluene (99%, Sinopharm Chemical Reagent Co., Ltd.), hexane ethanol (99%, Sinopharm Chemical Reagent Co., Ltd.), hexane (99%, Sigma Aldrich Co., St. Louis, MO, USA), benzene (99%, Sigma Aldrich Co.), trichloromethane (99%, Sinopharm Chemical Reagent Co., Ltd.), ferrocene (99.8%, Beijing Bailing Wei Technology Co., Ltd., Beijing, China), Octaethylnickel porphyrin (99.8%, Beijing Bailing Wei Technology Co., Ltd.), tetrabutylammonium hexafluorophosphate (electrochemical-grade, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), 1,2-dichlorobenzene (99%, Shanghai Aladdin Biochemical Technology Co., Ltd.), and graphite rods and graphite powder (Shanghai Fengyi Carbon Co., Shanghai, China).
Synthesis of Th@C1(17418)-C76. Th@C1(17418)-C76 was synthesized using the direct-current arc discharge method. ThO2 powders and graphite powder were mixed in an atomic molar ratio of Th:C = 1:24 (equivalent to a mass ratio of ThO2:C = 1:1.09). The resulting mixture was then uniformly filled into a hollow graphite rod and compacted to achieve dense packing. The filled graphite rod was placed into a tube furnace and subjected to a 12 h reduction process at 1000 °C under a nitrogen atmosphere, aiming to convert the metal oxide into a metal carbide.
After the reduction process, the compaction graphite rod containing the reduced material was utilized as the anode, while a solid graphite rod served as the cathode within a vacuum arc furnace. The furnace was evacuated to a pressure of 4 Pa and then filled with 200 Torr of helium gas. The distance between the electrodes was maintained at approximately 1 cm, and the current was set to 90 A using an electric welding machine. Plasma evaporation took place during the arc discharge process. The anode was manually rotated with the left hand to achieve the optimal evaporation positioning, while the cathode was rotated with the right hand to push the solid graphite rod towards the consumed mixture graphite rod. When the mixture rod was nearly depleted, the cathode was retracted rapidly to prevent any damage to the metallic target on the anode. Subsequently, the power supply of the electric welding machine was switched off. Once the furnace temperature cooled down to room temperature, air was introduced into the furnace to reach atmospheric pressure. The furnace cover was opened, and the carbon ash accumulated on the furnace wall was collected using tools like a test tube brush. The collected carbon ash was soaked in a solution of carbon disulfide and left to stand for 12 h. Once the first immersion was completed, the obtained carbon ash was immersed twice with the same operation, so that as much of the obtained fullerene sample as possible was extracted. The extraction of fullerene samples needs to be carried out in a glove box with specific filtration devices.
The carbon disulfide solution was separated from the solid carbon residues using a vacuum filtration device equipped with a 0.45 μm organic membrane, resulting in a carbon disulfide filtrate. The carbon disulfide solvent was then evaporated using rotary evaporation until no liquid remained, and an appropriate quantity of toluene solution was added to dissolve the residue. The resulting reddish brown solution underwent another round of filtration to eliminate the solid particles of fullerenes, resulting in a saturated crude extract solution of fullerenes, which could be further utilized for High-Performance Liquid Chromatography (HPLC) separation in subsequent steps.
Isolation of Th@C1(17418)-C76. The separation and purification of Th@C1(17418)-C76 was achieved using multi-stage HPLC procedures. The HPLC was performed with toluene as the mobile phase and a UV detector with a detection wavelength of 310 nm. A 10 mL chromatographic injector was used to inject samples of fullerenes dissolved in toluene into a preparative HPLC system for separation. Prior to injection, a disposable syringe filter was employed to filter the fullerene sample, mitigating the interference caused by solid particles during the chromatographic separation. Distinct chromatographic columns were employed for the separation at various stages of the process. Multiple HPLC columns, including a Buckyprep-M column (25 × 250 mm, Cosmosil, Nacalai Tesque Inc., Kyoto, Japan), a Buckprep column (10 × 250 mm, Cosmosil, Nacalai Tesque, Kyoto, Japan), and a 5PBB column (10 × 250 mm, Cosmosil, Nacalai Tesque, Kyoto, Japan), were utilized in the procedures. The purity of the target fractions was monitored by using MALDI-TOF during the separation process. In our mass spectrometry analysis, numerous signals corresponding to monometallic fullerenes based on thorium (Th), including Th@C76, Th@C80, Th@C82, Th@C84, Th@C86, and Th@C88, were observed. Among them, in particular, the isomer of Th@C76 reported in this study has not been experimentally documented before. Therefore, in subsequent separation steps, our focus was solely on the purification and characterization of Th@C76.
Th@C1(17418)-C76 was subjected to purification utilizing a four-stage high-performance liquid chromatography (HPLC) process, as shown in Figure S1. Initially, the compound was passed through a Buckyprep-Mcolumn (25 mm × 250 mm, Cosmosil, Nacalai Tesque) at a flow rate of 10 mL/min during the first stage of separation. Subsequently, the fraction eluted between 26 and 27.5 min (indicated by the color blue) was collected and reinjected into the Buckyprep column (10 mm × 250 mm, Cosmosil, Nacalai Tesque) for the second stage of separation, utilizing toluene as the eluent with a flow rate of 4 mL/min. The fraction containing Th@C1(17418)-C76 was obtained between 46 and 49.5 min (highlighted in red) and isolated accordingly. Moving on to the third stage of the separation, the 5PBB column (10 mm × 250 mm, Cosmosil, Nacalai Tesque) was employed, delivering a flow rate of 4 mL/min. The desired fraction, marked in orange, appeared within the 65 to 73 min range and specifically contained Th@C1(17418)-C76. Finally, the last stage involved passing the sample through the Buckyprep column (10 mm × 250 mm, Cosmosil, Nacalai Tesque) at a flow rate of 4 mL/min, bringing forth the achievement of pure Th@C1(17418)-C76. Notably, the confirmation of the extracted Th@C1(17418)-C76′s purity was impeccably established through the utilization of MALDI-TOF in positively charged mode, as visually exemplified in Figure 1.
Spectroscopic Studies of Th@C1(17418)-C76. The positive-ion mode matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) (Bruker, Karlsruhe, Germany) was employed for the mass characterization. A rotary evaporator was used to spin dry the fullerene sample under investigation. The substance was dissolved in carbon disulfide and the sample was transferred onto a mass spectrometry target using a pipette. A mass spectrometry analysis proceeded after the solvent evaporated. The purified Th@C1(17418)-C76 fullerene sample was dissolved in CS2 and the carbon disulfide solution of the fullerene was transferred into a 2 mL quartz cuvette. A control experiment was prepared by adding blank carbon disulfide to another quartz cuvette. The Cary 5000 UV-vis-NIR spectrophotometer (Agilent, Santa Clara, CA, USA) was used to measure the ultraviolet-visible-near-infrared absorption wavelengths of the sample in the detection range of 400–1600 nm. After completing the measurements, the carbon disulfide solution was collected from the sample in a glass vial and the quartz cuvette was rinsed with clean carbon disulfide solution. The Raman spectra were obtained using a Horiba Lab RAM HR Evolution Raman spectrometer using a laser at 633 nm.
Electrochemical Studies of Th@C1(17418)-C76. The cyclic voltammetry was performed using a conventional three-electrode system consisting of a glassy carbon electrode as the working electrode, a silver wire as the reference electrode, and a platinum wire as the counter electrode. A 0.05 M (n-Bu)4NPF6 solution was prepared by adding ortho-dichlorobenzene to the electrolyte, which contained tetrabutylammonium hexafluorophosphate. Subsequently, the purified sample of monometallic fullerene was dissolved in the solution for a cyclic voltammetry analysis using a CHI-660E electrochemical workstation. The reference material used for calibration was ferrocene, and the scan rate was set at 100 mV/s.
X-ray Crystallographic Study. Single crystals of Th@C76 were obtained, employing a solvent diffusion method characterized in a gradual process. A supersaturated solution of benzene (0.2 mL), encompassing NiII(OEP) (0.18 mg), was meticulously diffused into an impeccably concocted carbon disulfide solution (0.1 mL) of Th@C1(17418)-C76 (0.1 mg). Prior to the crystal growth procedure, the fabricated fullerene sample underwent a comprehensive purification and cleaning step utilizing n-hexane. Filtering and cleaning devices were obtained by using a rubber-tipped dropper. The filter paper was folded to obtain a filter core approximately 1 cm in size, precisely filling the neck of the pipette. The fullerene sample requiring filtration and cleaning was dissolved in carbon disulfide, followed by drying with nitrogen gas. A small amount of hexane was added, and ultrasonication was applied to suspend the fullerene sample in the hexane solution. Using a pipette, the mixture of hexane and fullerene samples was transferred to a prepared filtration pipette, with the rate of filtration controlled at one drop per second. As fullerene samples are insoluble in hexane, they remained on the filter core. The filter core was then rinsed with carbon disulfide, and the cleaning solution was collected, resulting in a fully cleaned fullerene sample.
In explicit detail, the carbon disulfide solution of Th@C1(17418)-C76 and the benzene solution of NiII(OEP) were introduced into a precisely maintained 5 mm inner diameter nuclear magnetic resonance (NMR) tube, adhering to an exact volumetric ratio of 1:2. Prior to crystal growth, the NMR tube necessitated a series of sequential cleansing steps. The NMR tube was thoroughly filled with solutions of soap, ethanol, n-hexane, and carbon disulfide in a successive manner. The subsequent cleansing process involved subjecting the NMR tube to ultrasonic irradiation for a duration exceeding 30 min, with three cycles of ultrasonic treatment for each solution. To ensure a meticulously controlled and hermetically sealed environment, the aperture of the NMR tube was securely enclosed using a specialized sealing film. This meticulously prepared assembly was subsequently transferred to a refrigerated environment, maintaining a constant temperature of 4 °C, and left undisturbed, allowing for a designated duration of one month.
After the designated duration, we used an optical microscope to observe the NMR tubes, black prism-shaped crystals, exhibiting a remarkable specular gleam, precipitated within the NMR tube. The meticulous selection of appropriately sized and shaped black prism-shaped crystals was conducted to facilitate a subsequent single-crystal X-ray diffraction analysis. These selected crystals were subjected to thorough testing and characterization employing a cutting-edge single-crystal diffractometer; the model was the Bruker D8 Venture with metaljet. The single-crystal X-ray data of Th@C1(17418)-C76 were collected at 120 K, respectively, on a diffractometer (APEX II; Bruker Analytik GmbH, Karlsruhe, Germany) equipped with a CCD collector. The multiscan method was used for absorption correction. The structures were solved using direct methods and refined on F2 using full-matrix least-squares using the SHELXL2014 crystallographic software packages. Hydrogen atoms were inserted at calculated positions and constrained with isotropic thermal parameters.
Crystal data for Th@C1(17418)-C76·[NiII(OEP)]·2C6H6: Mr = 1892.47, 0.1 mm × 0.08 mm × 0.06 mm, monoclinic, C2/c (No. 15), a = 44.778(2) Å, b = 14.9844(11) Å, c = 25.431(22) Å, α = 90°, β = 121.405(3)°, γ = 90°, V = 14564(22) Å3, Z = 8, ρcalcd = 1.756 g cm−3, μ(Ga Kα) = 6.129 mm−1, θ = 2.756–58.099, T = 120.0 K, R1 = 0.1339, wR2 = 0.2547 for all data; R1 = 0.0901, wR2 = 00.2256 for 10155 reflections (I > 2.0σ(I)) with 1220 parameters. Goodness-of-fit indicator1.074. The Maximum residual electron density was 1.429 e Å−3.

4. Conclusions

In conclusion, a novel non-IPR isomer of thorium-based mono-EMFs, the synthesis and characterization of Th@C1(17418)-C76 were accomplished through mass spectroscopy, single-crystal X-ray diffraction, Raman spectroscopy, and UV-vis-NIR spectroscopy. The crystallographic characterization unambiguously demonstrated that this isomer possesses a pair of adjacent pentagons, sharing the same carbon cage as the one previously reported for U@C1(17418)-C76. The major metal sites in U@C1(17418)-C76 and Th@C1(17418)-C76 are all situated at the intersection of two adjacent pentagons. Currently, the stabilization of the C1(17418)-C76 cage is only achieved by the encapsulation of U and Th, suggesting that the unique 5f electrons present in actinide metals enable them to stabilize uncommon carbon cage structures. In addition, despite similarity of their Raman vibrational spectra, the UV-vis-NIR absorption spectra of U@C1(17418)-C76 and Th@C1(17418)-C76 exhibit notable differences. This work introduces a novel addition to the group of thorium-based mono-EMFs, enhancing the comprehension of the distinctive carbon cage structure, electronic properties, and interactions exhibited by actinium-based EMFs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics11110422/s1, Figure S1: HPLC isolation procedures of Th@C1(17418)-C76 and the pure sample’s MALDI-TOF mass spectrum.; Figure S2: Ball and stick representation of disordered Th sites inside Th@C1(17418)-C76, five positions (Th1, Th2, Th3, Th4, and Th5) are observed.; Figure S3: Cyclic voltammograms of Th@C1(17418)-C76 in o-dichlorobenzene.; Table S1: Occupancies of disordered thorium sites in Th@C1(17418)-C76.; Table S2: Closest Th-Cage distances (Å) in Th@C1(17418)-C76.; Table S3: Crystallographic information of Th@C1(17418)-C76.

Author Contributions

Conceptualization, Y.X.; funding acquisition, N.C.; investigation, Y.X., Y.S. and Y.-R.Y.; project administration, N.C. and Q.M.; supervision, N.C.; writing—original draft, Y.X.; writing—review and editing, N.C. and Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Foundation China (NSFC 52172051 and 22301288) and the Anhui Provincial Natural Science Foundation (2308085MB31).

Data Availability Statement

The crystallographic data can be obtained free of charge from the Cambridge Crystallographic Data Centre with the CCDC numbers 2295545 via www.ccdc.cam.ac.uk/data_request/cif (accessed on 18 September 2023). The other data are available upon reasonable request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  2. Yang, S.; Wei, T.; Jin, F. When metal clusters meet carbon cages: Endohedral clusterfullerenes. Chem. Soc. Rev. 2017, 46, 5005–5058. [Google Scholar] [CrossRef]
  3. Bao, L.; Peng, P.; Lu, X. Bonding inside and outside Fullerene Cages. Acc. Chem. Res. 2018, 51, 810–815. [Google Scholar] [CrossRef] [PubMed]
  4. Chai, Y.; Guo, T.; Jin, C.M.; Haufler, R.E.; Chibante, L.P.F.; Fure, J.; Wang, L.H.; Alford, J.M.; Smalley, R.E. Fullerenes with Metals Inside. J. Phys. Chem. 1991, 95, 7564–7568. [Google Scholar] [CrossRef]
  5. Slanina, Z.; Nagase, S. Sc3N@C80: Computations on the Two-Isomer Equilibrium at High Temperatures. Chem. Phys. Chem. 2005, 6, 2060–2063. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, T.-S.; Feng, L.; Wu, J.-Y.; Xu, W.; Xiang, J.-F.; Tan, K.; Ma, Y.-H.; Zheng, J.-P.; Jiang, L.; Lu, X.; et al. Planar Quinary Cluster inside a Fullerene Cage: Synthesis and Structural Characterizations of Sc3NC@C80-Ih. J. Am. Chem. Soc. 2010, 132, 16362–16364. [Google Scholar] [CrossRef] [PubMed]
  7. Kodama, T.; Fujii, R.; Miyake, Y.; Sakaguchi, K.; Nishikawa, H.; Ikemoto, I.; Kikuchi, K.; Achiba, Y. Structural study of four Ca@C82 isomers by 13C NMR spectroscopy. Chem. Phys. Lett. 2003, 377, 197–200. [Google Scholar] [CrossRef]
  8. Xu, Z.; Nakane, T.; Shinohara, H. Production and Isolation of Ca@C82 (I–IV) and Ca@C84 (I,II) Metallofullerenes. J. Am. Chem. Soc. 1996, 118, 11309–11310. [Google Scholar] [CrossRef]
  9. Xu, W.; Feng, L.; Calvaresi, M.; Liu, J.; Liu, Y.; Niu, B.; Shi, Z.; Lian, Y.; Zerbetto, F. An experimentally observed trimetallofullerene Sm3@Ih-C80: Encapsulation of three metal atoms in a cage without a nonmetallic mediator. J. Am. Chem. Soc. 2013, 135, 4187–4190. [Google Scholar] [CrossRef]
  10. Kurihara, H.; Lu, X.; Iiduka, Y.; Mizorogi, N.; Slanina, Z.; Tsuchiya, T.; Akasaka, T.; Nagase, S. Sc2C2@C80 Rather than Sc2@C82: Templated Formation of Unexpected C2v(5)-C80 and Temperature-Dependent Dynamic Motion of Internal Sc2C2 Cluster. J. Am. Chem. Soc. 2011, 133, 2382–2385. [Google Scholar] [CrossRef]
  11. Hu, S.; Liu, T.; Shen, W.; Slanina, Z.; Akasaka, T.; Xie, Y.; Uhlik, F.; Huang, W.; Lu, X. Isolation and Structural Characterization of Er@C2v(9)-C82 and Er@Cs(6)-C82: Regioselective Dimerization of a Pristine Endohedral Metallofullerene Induced by Cage Symmetry. Inorg. Chem. 2019, 58, 2177–2182. [Google Scholar] [CrossRef] [PubMed]
  12. Shen, W.; Bao, L.; Hu, S.; Yang, L.; Jin, P.; Xie, Y.; Akasaka, T.; Lu, X. Crystallographic characterization of Lu2C2n (2n = 76–90): Cluster selection by cage size. Chem. Sci. 2019, 10, 829–836. [Google Scholar] [CrossRef] [PubMed]
  13. Shen, W.; Bao, L.; Wu, Y.; Pan, C.; Zhao, S.; Fang, H.; Xie, Y.; Jin, P.; Peng, P.; Li, F.-F.; et al. Lu2@C2n (2n = 82, 84, 86): Crystallographic Evidence of Direct Lu–Lu Bonding between Two Divalent Lutetium Ions Inside Fullerene Cages. J. Am. Chem. Soc. 2017, 139, 9979–9984. [Google Scholar] [CrossRef] [PubMed]
  14. Ross, R.B.; Cardona, C.M.; Guldi, D.M.; Sankaranarayanan, S.G.; Reese, M.O.; Kopidakis, N.; Peet, J.; Walker, B.; Bazan, G.C.; Van Keuren, E.; et al. Endohedral fullerenes for organic photovoltaic devices. Nat. Mater. 2009, 8, 208–212. [Google Scholar] [CrossRef]
  15. Liu, Y.; Chen, C.; Qian, P.; Lu, X.; Sun, B.; Zhang, X.; Wang, L.; Gao, X.; Li, H.; Chen, Z.; et al. Gd-metallofullerenol nanomaterial as non-toxic breast cancer stem cell-specific inhibitor. Nat. Commun. 2015, 6, 5988. [Google Scholar] [CrossRef]
  16. Zhang, K.; Wang, C.; Zhang, M.; Bai, Z.; Xie, F.F.; Tan, Y.Z.; Guo, Y.; Hu, K.J.; Cao, L.; Zhang, S.; et al. A Gd@C82 single-molecule electret. Nat. Nanotech. 2020, 15, 1019–1024. [Google Scholar] [CrossRef] [PubMed]
  17. Okamura, N.; Yoshida, K.; Sakata, S.; Hirakawa, K. Electron transport in endohedral metallofullerene Ce@C82 single-molecule transistors. Appl. Phys. Lett. 2015, 106, 043108. [Google Scholar] [CrossRef]
  18. Wakahara, T.; Kobayashi, J.; Yamada, M.; Maeda, Y.; Tsuchiya, T.; Okamura, M.; Akasaka, T.; Waelchli, M.; Kobayashi, K.; Nagase, S.; et al. Characterization of Ce@C82 and its anion. J. Am. Chem. Soc. 2004, 126, 4883–4887. [Google Scholar] [CrossRef]
  19. Akasaka, T.; Okubo, S.; Kondo, M.; Maeda, Y.; Wakahara, T.; Kato, T.; Suzuki, T.; Yamamoto, K.; Kobayashi, K.; Nagase, S. Isolation and characterization of two Pr@C82 isomers. Chem. Phys. Lett. 2000, 319, 153–156. [Google Scholar] [CrossRef]
  20. Ding, J.; Lin, N.; Weng, L.-T.; Cue, N.; Yang, S. Isolation and characterization of a new metallofullerene Nd@C82. Chem. Phys. Lett. 1996, 261, 92–97. [Google Scholar] [CrossRef]
  21. Hu, Z.; Hao, Y.; Slanina, Z.; Gu, Z.; Shi, Z.; Uhlík, F.; Zhao, Y.; Feng, L. Popular C82 Fullerene Cage Encapsulating a Divalent Metal Ion Sm2+: Structure and Electrochemistry. Inorg. Chem. 2015, 54, 2103–2108. [Google Scholar] [CrossRef] [PubMed]
  22. Sun, B.Y.; Sugai, T.; Nishibori, E.; Iwata, K.; Sakata, M.; Takata, M.; Shinohara, H. An Anomalous Endohedral Structure of Eu@C82 Metallofullerenes. Angew. Chem. Int. Ed. 2005, 44, 4568–4571. [Google Scholar] [CrossRef] [PubMed]
  23. Senapati, L.; Schrier, J.; Whaley, K.B. Electronic Transport, Structure, and Energetics of Endohedral Gd@C82 Metallofullerenes. Nano Lett. 2004, 4, 2073–2078. [Google Scholar] [CrossRef]
  24. Iwasaki, K.; Wanita, N.; Hino, S.; Yoshimura, D.; Okazaki, T.; Shinohara, H. Ultraviolet photoelectron spectra of Tb@C82. Chem. Phys. Lett. 2004, 398, 389–392. [Google Scholar] [CrossRef]
  25. Iida, S.; Kubozono, Y.; Slovokhotov, Y.; Takabayashi, Y.; Kanbara, T.; Fukunaga, T.; Fujiki, S.; Emura, S.; Kashino, S. Structure and electronic properties of Dy@C82 studied by UV–VIS absorption, X-ray powder diffraction and XAFS. Chem. Phys. Lett. 2001, 338, 21–28. [Google Scholar] [CrossRef]
  26. Huang, H.J.; Yang, S.H.; Zhang, X.X. Magnetic Behavior of Pure Endohedral Metallofullerene Ho@C82: A Comparison with Gd@C82. J. Phys. Chem. B 1999, 103, 5928–5932. [Google Scholar] [CrossRef]
  27. Sanakis, Y.; Tagmatarchis, N.; Aslanis, E.; Ioannidis, N.; Petrouleas, V.; Shinohara, H.; Prassides, K. Dual-Mode X-Band EPR Study of Two Isomers of the Endohedral Metallofullerene Er@C82. J. Am. Chem. Soc. 2001, 123, 9924–9925. [Google Scholar] [CrossRef]
  28. Suzuki, M.; Slanina, Z.; Mizorogi, N.; Lu, X.; Nagase, S.; Olmstead, M.M.; Balch, A.L.; Akasaka, T. Single-Crystal X-ray Diffraction Study of Three Yb@C82 Isomers Cocrystallized with NiII(octaethylporphyrin). J. Am. Chem. Soc. 2012, 134, 18772–18778. [Google Scholar] [CrossRef]
  29. Iwamoto, M.; Ogawa, D.; Yasutake, Y.; Azuma, Y.; Umemoto, H.; Ohashi, K.; Izumi, N.; Shinohara, H.; Majima, Y. Molecular Orientation of Individual Lu@C82 Molecules Demonstrated by Scanning Tunneling Microscopy. J. Phys. Chem. C 2010, 114, 14704–14709. [Google Scholar] [CrossRef]
  30. Wang, Y.F.; Morales-Martinez, R.; Zhang, X.X.; Yang, W.; Wang, Y.X.; Rodriguez-Fortea, A.; Poblet, J.M.; Feng, L.; Wang, S.; Chen, N. Unique Four-Electron Metal-to-Cage Charge Transfer of Th to a C82 Fullerene Cage: Complete Structural Characterization of Th@C3v(8)-C82. J. Am. Chem. Soc. 2017, 139, 5110–5116. [Google Scholar] [CrossRef]
  31. Jin, M.; Zhuang, J.; Wang, Y.; Yang, W.; Liu, X.; Chen, N. Th@Td(19151)-C76: A Highly Symmetric Fullerene Cage Stabilized by a Tetravalent Actinide Metal Ion. Inorg. Chem. 2019, 58, 16722–16726. [Google Scholar] [CrossRef] [PubMed]
  32. Yan, Y.; Morales-Martinez, R.; Zhuang, J.; Yao, Y.R.; Li, X.; Poblet, J.M.; Rodriguez-Fortea, A.; Chen, N. Th@D5h(6)-C80: A highly symmetric fullerene cage stabilized by a single metal ion. Chem. Commun. 2021, 57, 6624–6627. [Google Scholar] [CrossRef] [PubMed]
  33. Meng, Q.; Morales-Martinez, R.; Zhuang, J.; Yao, Y.R.; Wang, Y.; Feng, L.; Poblet, J.M.; Rodriguez-Fortea, A.; Chen, N. Synthesis and Characterization of Two Isomers of Th@C82: Th@C2v(9)-C82 and Th@C2(5)-C82. Inorg. Chem. 2021, 60, 11496–11502. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, Y.; Morales-Martinez, R.; Cai, W.; Zhuang, J.; Yang, W.; Echegoyen, L.; Poblet, J.M.; Rodriguez-Fortea, A.; Chen, N. Th@C1(11)-C86: An actinide encapsulated in an unexpected C86 fullerene cage. Chem. Commun. 2019, 55, 9271–9274. [Google Scholar] [CrossRef]
  35. Cai, W.; Morales-Martínez, R.; Zhang, X.; Najera, D.; Romero, E.L.; Metta-Magaña, A.; Rodríguez-Fortea, A.; Fortier, S.; Chen, N.; Poblet, J.M.; et al. Single crystal structures and theoretical calculations of uranium endohedral metallofullerenes (U@C2n, 2n = 74, 82) show cage isomer dependent oxidation states for U. Chem. Sci. 2017, 8, 5282–5290. [Google Scholar] [CrossRef] [PubMed]
  36. Cai, W.; Abella, L.; Zhuang, J.; Zhang, X.; Feng, L.; Wang, Y.; Morales-Martínez, R.; Esper, R.; Boero, M.; Metta-Magaña, A.; et al. Synthesis and Characterization of Non-Isolated-Pentagon-Rule Actinide Endohedral Metallofullerenes U@C1(17418)-C76, U@C1(28324)-C80, and Th@C1(28324)-C80: Low-Symmetry Cage Selection Directed by a Tetravalent Ion. J. Am. Chem. Soc. 2018, 140, 18039–18050. [Google Scholar] [CrossRef]
  37. Jin, P.; Liu, C.; Li, Y.; Li, L.; Zhao, Y. Th@C76. Computational characterization of larger actinide endohedral fullerenes. Int. J. Quantum. Chem. 2018, 118, e25501. [Google Scholar] [CrossRef]
  38. Zhao, P.; Zhao, X.; Ehara, M. Theoretical Insights into Monometallofullerene Th@C76: Strong Covalent Interaction between Thorium and the Carbon Cage. Inorg. Chem. 2018, 57, 2961–2964. [Google Scholar] [CrossRef]
  39. Bao, L.; Li, Y.; Yu, P.; Shen, W.; Jin, P.; Lu, X. Preferential Formation of Mono-Metallofullerenes Governed by the Encapsulation Energy of the Metal Elements: A Case Study on Eu@C2n (2n = 74–84) Revealing a General Rule. Angew. Chem. Int. Ed. 2020, 59, 5259–5262. [Google Scholar] [CrossRef]
  40. Hao, Y.; Feng, L.; Xu, W.; Gu, Z.; Hu, Z.; Shi, Z.; Slanina, Z.; Uhlik, F. Sm@C2v(19138)-C76: A Non-IPR Cage Stabilized by a Divalent Metal Ion. Inorg. Chem. 2015, 54, 4243–4248. [Google Scholar] [CrossRef]
Figure 1. HPLC chromatogram of purified Th@C76 on a Buckyprep column using toluene as the eluent, with a 4 mL/min flow rate. The mass spectra of Th@C76 and associated empirical and hypothetical isotopic profiles of the sample are shown in the inset.
Figure 1. HPLC chromatogram of purified Th@C76 on a Buckyprep column using toluene as the eluent, with a 4 mL/min flow rate. The mass spectra of Th@C76 and associated empirical and hypothetical isotopic profiles of the sample are shown in the inset.
Inorganics 11 00422 g001
Figure 2. ORTEP drawings showing the relationship between the EMF and porphyrin moieties for Th@C1(17418)-C76·[NiII(OEP)] with 15% thermal ellipsoids. Only the major cage orientation and predominant thorium sites are shown.
Figure 2. ORTEP drawings showing the relationship between the EMF and porphyrin moieties for Th@C1(17418)-C76·[NiII(OEP)] with 15% thermal ellipsoids. Only the major cage orientation and predominant thorium sites are shown.
Inorganics 11 00422 g002
Figure 3. Perspective drawings showing the interaction of the major thorium closest cage portion in Th@C1(17418)-C76.
Figure 3. Perspective drawings showing the interaction of the major thorium closest cage portion in Th@C1(17418)-C76.
Inorganics 11 00422 g003
Figure 4. UV-vis-NIR absorption spectra of Th@C1(17418)-C76 and (a) U@C1(17418)-C76 and (b) Th@Td(19151)-C76 in CS2.
Figure 4. UV-vis-NIR absorption spectra of Th@C1(17418)-C76 and (a) U@C1(17418)-C76 and (b) Th@Td(19151)-C76 in CS2.
Inorganics 11 00422 g004
Figure 5. Low-energy Raman spectra of Th@C1(17418)-C76 at 633 nm excitation.
Figure 5. Low-energy Raman spectra of Th@C1(17418)-C76 at 633 nm excitation.
Inorganics 11 00422 g005
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

Xia, Y.; Shen, Y.; Yao, Y.-R.; Meng, Q.; Chen, N. Synthesis and Characterization of a Novel Non-Isolated-Pentagon-Rule Isomer of Th@C76:Th@C1(17418)-C76. Inorganics 2023, 11, 422. https://doi.org/10.3390/inorganics11110422

AMA Style

Xia Y, Shen Y, Yao Y-R, Meng Q, Chen N. Synthesis and Characterization of a Novel Non-Isolated-Pentagon-Rule Isomer of Th@C76:Th@C1(17418)-C76. Inorganics. 2023; 11(11):422. https://doi.org/10.3390/inorganics11110422

Chicago/Turabian Style

Xia, Yunpeng, Yi Shen, Yang-Rong Yao, Qingyu Meng, and Ning Chen. 2023. "Synthesis and Characterization of a Novel Non-Isolated-Pentagon-Rule Isomer of Th@C76:Th@C1(17418)-C76" Inorganics 11, no. 11: 422. https://doi.org/10.3390/inorganics11110422

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