• Matter and Radiation at Extremes
  • Vol. 4, Issue 2, 28201 (2019)
Cuiying Pei and Lin Wang*
Author Affiliations
  • Center for High Pressure Science and Technology Advanced Research (HPSTAR), 1690 Cailun Rd., Pudong District, Shanghai 201203, People’s Republic of China
  • show less
    DOI: 10.1063/1.5086310 Cite this Article
    Cuiying Pei, Lin Wang. Recent progress on high-pressure and high-temperature studies of fullerenes and related materials[J]. Matter and Radiation at Extremes, 2019, 4(2): 28201 Copy Citation Text show less

    Abstract

    Polymerization of fullerenes is an interesting topic that has been studied for almost three decades. A rich polymeric phase diagram of C60 has been drawn under a variety of pressure P and temperature T conditions. Knowledge of the targeted preparation and structural control of fullerene polymers has become increasingly important because of their utility in producing novel fullerene-based architectures with unusual properties and potential applications. This paper focuses on the polymeric phases of fullerenes and their derivatives under high P and/or high T. First, the polymerization behavior and the various polymeric phases of C60 and C70 under such conditions are briefly reviewed. A summary of the polymerization process of intercalated fullerenes is then presented, and a synthetic strategy for novel structural and functional fullerene polymers is proposed. Finally, on the basis of the results of recent research, a proposal is made for further studies of endohedral fullerenes at high P.

    I. INTRODUCTION

    Pressure is a fundamental thermodynamic variable that can dramatically alter the atomic and electronic structures of materials, resulting in concomitant changes to their properties.1 The C=C bonds in fullerenes allow them to form covalent intermolecular bonds when treated at high pressure and high temperature (HPHT). The preparation path based on the reaction diagram and the treatment time greatly influences the completeness of transformation and the degree of polymerization. In addition, the hydrostatic nature of the compression environment plays an important role in bonding orientation and polymerization ordering. As a result, dimers, ordered one-dimensional (1D), 2D, and 3D polymers, and disordered graphite-like, diamond-like phases, etc., have been reported with a variety of novel polymerized fullerene architectures and a range of unusual properties and potential applications.2–11

    The controllable polymerization of fullerenes to create new materials by using molecular confinement or co-intercalation by template molecules in the reactant has been extensively studied under a wide range of conditions of pressure P and temperature T. Solvated fullerenes, in which solvent molecules separate C60 and C70 molecules, are a series of crystalline materials with high stability, tunable metrics, and functionality. The intercalated solvent molecules play an important role in the construction of solvate lattices with different structures and further influence the phase transition at high P.12–17 In addition, several metastable doped fullerene polymers have been synthesized under HPHT, such as C60/Ni(OEP),18,19 C60/CNT,20 and Na4C60.21 The unique initial lattice structures and the different degrees of spatial confinement provided by the dopants lead to different polymeric phases accompanied by distinctive physical and electrical properties.

    In 2001, Buga et al.22 reported the P- and T-induced polymerization of the mono-atom endohedral fullerene La@C82 for the first time. Later, Liu and co-workers studied the structural transformations of Sm@C88 (Ref. 23) and solvated Sm@C90 (Ref. 24). The HOMO–LUMO gap decreased with increasing P until the carbon cage deformed, but no phase transition occurred. Pei et al.25 observed for the first time that Lu3N@C80 underwent a reversible n- to p-type inversion under high pressure, with the p-type semiconductor being stable up to 25 GPa. The role of the intercage metal atom is to restrain the compression of the adjacent bonds and protect the endohedral fullerene against collapse, as well as decreasing and postponing the change in band gap. To date, however, the structure and properties of endohedral fullerenes under high P remain undefined.

    This paper focuses on the properties and phases of fullerenes and their derivatives under HPHT. Four topics will be discussed: (i) how fullerenes are covalently incorporated into a polymer phase under HPHT; (ii) the correlation between the structural evolution and the novel properties of the polymeric phase; (iii) the effect of pressure on the polymerization behavior of doped or intercalated fullerenes; (iv) the polymeric fullerenes and their derivatives as new and versatile building blocks to prepare a wide variety of new fullerene-based materials governed by weak intermolecular interactions, showing unprecedented architectures and a wide range of potential applications.

    II. HPHT RESEARCH ON C60

    To date, many different strategies have been exploited to prepare polymeric fullerenes. Chemical reactions can link C60 cages either covalently or supramolecularly, and the polymeric phases exhibit outstanding structural, electrochemical, and photophysical properties.26 Photopolymerization,27 pressure- and/or temperature-induced polymerization,2–4 charge-transfer polymerization mediated by metals,28 electron-beam-induced polymerization,29 and plasma-induced polymerization30,31 can be used to prepare novel polymeric fullerenes. In this section, we will present a general “panoramic” view of all-carbon polymeric fullerenes under high P and/or high T. Since the discovery of polymeric fullerene synthesized at high P in 1994,32 there has been intense interest in the characterization of the polymeric phase products and in clarification of the mechanisms underlying the process of polymerization, because of the potentially useful physical, biological, and chemical properties of these products.33,34 In particular, there have been extensive investigations of the influence of P and T treatment on polymerization,35 on infrared spectroscopy (IR) and Raman vibrational properties,36 and on photoluminescence37 and magnetism.38

    P-induced polymerization under various T and P has been studied extensively in recent years.2 When moderate hydrostatic or quasi-hydrostatic P at high T is applied to pristine C60, irreversible polymerization occurs over a limited range of T and P. As shown in the P vs. T phase diagram (Fig. 1), C60 polymers have orthorhombic, tetragonal, rhombohedral, or cross-linked phases. Recently, single crystals of these polymeric phases have been successfully prepared (the orthorhombic phase in Ref. 39, the tetragonal phase in Refs. 40–43, and the rhombohedral phase in Ref. 44). Under ultrahigh P and T, 3D polymers have been discovered, but, as yet, without a uniform structure.

    Map of the P–T plane showing the various phases of C60 created under different conditions. Symbols denote some of the P–T coordinates where experiments have been carried out and samples have been treated for at least 1 min. (—) denotes the diamond/graphite equilibrium line, while (– – –) divides the semimetal and semiconductor states at T > 1100 K and the “soft” and “hard” states at T et al., Carbon 36, 319–343 (1998). Copyright 1998 Elsevier.

    Figure 1.Map of the P–T plane showing the various phases of C60 created under different conditions. Symbols denote some of the P–T coordinates where experiments have been carried out and samples have been treated for at least 1 min. (—) denotes the diamond/graphite equilibrium line, while (– – –) divides the semimetal and semiconductor states at T > 1100 K and the “soft” and “hard” states at T < 1100 K. Reprinted with permission from V. D. Blank et al., Carbon 36, 319–343 (1998). Copyright 1998 Elsevier.

    A. Bonding in HPHT polymeric C60

    C60, an electron-deficient polyene, tends to react with other molecules via radical, nucleophilic, or cycloaddition mechanisms. Fullerene can be functionalized via [2 + n] cycloaddition (n = 1, 2, 3, 4) reactions.45 In P-induced polymerization, a [2 + 2] cycloaddition reaction between two [6,6] double bonds of near-neighbor C60 cages is the favored reaction pathway (Fig. 2). A new cyclobutane ring is formed by covalent linkage in the polymeric product. On the other hand, the formation of C60 intermolecular bonds is a thermally activated process, and the polymerization reaction is very slow at room temperature. Davydov et al.46 have shown that the activation energy for the formation of the C60 dimer is approximately 135 kJ mol−1. However, at ambient P and T between 300 °C and 400 °C, the polymeric product is metastable and reverts to pristine C60.32 Hence, P is necessary for the formation of a stable polymeric product because it reduces the intermolecular distance sufficiently for chemical bond formation to take place.

    Formation of C60 polymers by [2 + 2] cycloaddition of pristine C60. (The figure has been redrawn based on Ref. 26.) Reprinted with permission from F. Giacalone and N. Martín, Chem. Rev. 106, 5136–5190 (2006). Copyright 2006 American Chemical Society.

    Figure 2.Formation of C60 polymers by [2 + 2] cycloaddition of pristine C60. (The figure has been redrawn based on Ref. 26.) Reprinted with permission from F. Giacalone and N. Martín, Chem. Rev. 106, 5136–5190 (2006). Copyright 2006 American Chemical Society.

    In the 3D C60 polymer, [3 + 3] cycloaddition bonds form between 2D layers to construct a 3D framework, and it is a fragment of the di-(1.5)(6,10)methanol-cyclodecane molecule.47,48 All the equivalent 3 and 6 atoms of each C60 molecule are bonded in pairs with the corresponding atoms of the neighboring (x,y) layer molecules. The intermolecular bond lengths are 0.151 nm and 0.161 nm for the [3 + 3] and [2 + 2] cycles, respectively.47

    The distance between the center-of-mass positions of the two nearest-neighbor C60 in face-centered cubic (fcc) bulk C60 is 10.1 Å.49 Generally, it is 9.1 Å in C60 [2 + 2] cycloaddition polymers and about 9.4 Å when a single C–C bond links two cages.50,51 For example, the intermolecular distance along the C60 orthorhombic chains is 9.14 Å, as characterized by single-crystal X-ray diffraction experiments.39 In confined C60, the nearest-neighbor cage distance approaches a value of 8.45 Å near 10 GPa.52

    Alkali-metal-doped fullerene polymerizes under high T or HPHT. However, the framework is constructed using single carbon bonds.53 In the case of endohedral fullerenes with C80, [2 + 2] cycloaddition reactions occur at both the [6,6] and [5,6] bonds of the cage. The [6,6]-regioisomer is a kinetic product, while the [5,6]-regioisomer is a thermodynamic product.54

    B. Spectroscopy of monomeric, dimeric, and polymeric phases

    Polymerization changes both the microscopic and macroscopic properties of a material. Insolubility of the polymeric materials in those solvents that do dissolve C60 is direct proof of the polymerization. Polymerization is also confirmed by shifts in the vibrational modes and activation of Raman and infrared modes due to the reduced molecular symmetry as well as to the greater degree of freedom, and these also indicate the degree of polymerization (Fig. 3). According to the symmetry of the displacement patterns, there are 46 distinct intramolecular vibrations for C60. Group theory indicates ten first-order Raman-active modes, including two breathing Ag modes related to the symmetric oscillations of the entire molecule and pentagons together with eight Hg modes.55–60 In addition, four intramolecular T1u modes are infrared-active at first order.55,61,62 The Raman (Fig. 3) and infrared (IR) spectra of the pressure-dimerized state, as well as those of the orthorhombic, tetragonal, and rhombohedral pressure-polymerized phases of C60, have been investigated.55

    Raman spectra of pristine C60 (a), polymerized dimers (b), and orthorhombic phase (c), excited by a 1064 nm line, and of the orthorhombic (d), tetragonal (e), and rhombohedral (f) phases, excited by a 568.2 nm line. Reprinted with permission from V. A. Davydov et al., Phys. Rev. B 61, 11936 (2000). Copyright 2000 American Physical Society.

    Figure 3.Raman spectra of pristine C60 (a), polymerized dimers (b), and orthorhombic phase (c), excited by a 1064 nm line, and of the orthorhombic (d), tetragonal (e), and rhombohedral (f) phases, excited by a 568.2 nm line. Reprinted with permission from V. A. Davydov et al., Phys. Rev. B 61, 11936 (2000). Copyright 2000 American Physical Society.

    The number of covalent bonds per C60 molecule increases with increasing P and T. Accordingly, the formation of intermolecular bonds shifts the electron density away from the remaining double bonds, which become weaker. In particular, the Ag(2) mode is one of the most sensitive to polymer structure and charge transfer. As shown in Fig. 3, it downshifts from 1470 cm−1 in the monomer to 1465 cm−1 in the dimer, which corresponds to two sp3-like coordinated carbon atoms per fullerene molecule.63 Upon further polymerization, the softening is 12 cm−1 in the linear polymeric chains, 20 cm−1 in the planar tetragonal phase, and 60 cm−1 in the 2D rhombohedral polymers.55 Apart from this, the presence of sp3 intermolecular bonds leads to the appearance of new bonds in the range 900–1000 cm−1.

    Thermally induced dissociation of the polymeric bonds in C60 polymers can be monitored as a function of time at high T. For example, the band shape and the intensity of the Ag(2) mode change with increasing time.64 Novel polymeric C60 phases can also be recognized by Raman spectroscopy measurements. For instance, a branched-chain C60 polymer shows a characteristic mode at 1454 cm−1, a more pronounced splitting of Hg(1), and a new Raman-activated mode around Ag(1).63

    The symmetry of C60 is reduced from Ih to D2h in the orthorhombic and tetrahedral polymers and to D3d in the rhombohedral phase.65 Accordingly, the T1u(2) mode is primarily polarized in the stretched directions and dramatically downshifts more than 50 cm−1 in the IR spectra of the orthorhombic and rhombohedral polymers. The large distortions of the C60 sphere due to the numbers of cycloaddition connections demand a distorted polymer C60 ball with relaxed atomic positions.

    C. Dimerization of C60 at HPHT

    Before discussing the formation and structures of polymeric C60, we briefly review the structural features of C60 at ambient P. C60 molecules exhibit a rotationally disordered fcc structure above 260 K, free rotation at high T, and an orientationally disordered phase at 90 K. Under compression, the crystal undergoes an orientational transition at around 0.4 GPa and ambient T, which transforms into a simple cubic (sc) structure.2,13,66 During further shortening of the intermolecular distance, [6,6] double bonds open and four-membered rings link nearest neighbor cages. A solid-state mechanochemical method shows that the C4 ring connecting the cages is square rather than rectangular. As shown in Fig. 4, C60 dimers are disordered in position and orientation within an average cubic lattice derived from the monomer.67 C60 dimerization begins either within the fcc phase, in which the cage rotates freely, or within the sc structure with nearest neighbors in the P-DB (C=C double bond facing pentagon) or H-DB (C=C double bond facing hexagon) configuration.68 This mechanism can be observed even at room temperature at about 1.0 GPa.69 However, this thermally activated process is very slow at ambient T. As a result, direct synthesis of pure C60 dimers at HPHT is not yet possible. Typical conditions for synthesizing the C60 dimer are 1.5 GPa, 423 K, and 1000 s, with a yield of about 80%.55 When the C60 molecules are linked, the icosahedral symmetry is drastically reduced. As a consequence, the degenerate energy levels near the Fermi level are removed, and thus the optical band gap narrows.70

    A random distribution of C60 dimers (and isolated monomers) in the (001) plane of a model crystal, which accounts reasonably well for the experimentally observed diffuse scattering intensity distribution. Only in-plane and out-of-plane C60’s that form dimers with an in-plane C60 are shown. Reprinted with permission from R. Moret et al., Eur. Phys. J. B 37, 25–37 (2004). Copyright 2004 Springer Nature.

    Figure 4.A random distribution of C60 dimers (and isolated monomers) in the (001) plane of a model crystal, which accounts reasonably well for the experimentally observed diffuse scattering intensity distribution. Only in-plane and out-of-plane C60’s that form dimers with an in-plane C60 are shown. Reprinted with permission from R. Moret et al., Eur. Phys. J. B 37, 25–37 (2004). Copyright 2004 Springer Nature.

    D. 1D polymerization of C60 at HPHT

    At higher T (<700 K) and P (∼0.7–9 GPa), long linear chains with two four-membered rings per C60 cage form. Parallel straight chains along the original 〈110〉 directions in the ordered orthorhombic structure are observed in powder and single-crystal diffraction results, although disordered branched chains possibly exist as well (Fig. 5). At the same time, a large and rapid change in volume at P ∼ 1 GPa has been observed.71 The XRD pattern confirms that the nearest-neighbor distance (about 9.1–9.2 Å) in the orthorhombic phase polymer decreases by almost 10% compared with 10.02 Å in pristine C60, owing to the formation of covalent bonds.39,55 It is interesting to note that two different orthorhombic phase C60 polymers can be formed, corresponding to “high” (above ∼2–3 GPa) and “low” P regions. These structures are recognized by the different orientations of the chains around their axes. The chain links are parallel to the plane in the orthorhombic (O) structure,33 while they are alternately tilted by angles +μ and −μ from chain to chain in another orthorhombic (O′) phase.39,72 As a result, the O and O′ phases demonstrate a pseudo-tetragonal Immm space group and orthorhombic Pmnn space group, respectively. It is hard to estimate the angle μ, but lattice energy minimization predicts an optimal value of 29°.68

    Schematic views of the polymer chains running along 〈110〉 cubic directions of the parent monomer lattice, for the different polymer structures. The chains are normal to the figure. The orientation of the four-membered rings is represented by the shaded bars that form an angle μ with the 〈001〉 cubic directions. The broken lines represent the 2D tetragonal and rhombohedral polymer layers that can be obtained, at least schematically, by connecting the C60 molecules of the chains by supplementary four-membered rings. Reprinted with permission from R. Moret et al., AIP Conf. Proc. 544, 81–84 (2000). Copyright 2000 AIP Publishing LLC.

    Figure 5.Schematic views of the polymer chains running along 〈110〉 cubic directions of the parent monomer lattice, for the different polymer structures. The chains are normal to the figure. The orientation of the four-membered rings is represented by the shaded bars that form an angle μ with the 〈001〉 cubic directions. The broken lines represent the 2D tetragonal and rhombohedral polymer layers that can be obtained, at least schematically, by connecting the C60 molecules of the chains by supplementary four-membered rings. Reprinted with permission from R. Moret et al., AIP Conf. Proc. 544, 81–84 (2000). Copyright 2000 AIP Publishing LLC.

    E. 2D polymerization of C60 at HPHT

    At higher T (>700 K) and P (1.5–9 GPa), four or six four-membered rings per C60 molecule participate in the construction of the 2D C60 polymer. The formation of the polymer layers by cross-coupling of chains appears in two possible unit cells (tetragonal and trigonal). After that, the layers stack in a close-packed arrangement to form the corresponding tetragonal or rhombohedral structures. When [2 + 2] cycloaddition occurs in the (001) plane of the original fcc phase, a 2D tetragonal phase is formed [shown in Fig. 3(e)].2 For example, a single crystal of the “tetragonal” polymer was obtained when C60 was treated at 2.5 GPa and 500 °C.43 A single-crystal study identified the tetragonal phase to be a pseudo-tetragonal packing of translationally identical adjacent 2D layers with an orthorhombic space group Immm.43 In this case, the chain orientations are almost parallel. In another case, the chain links are alternately parallel and perpendicular to the stacking axis and exhibit a P42/mmc space group. Interestingly, Wågberg et al.74 reported that the structural transition from an orthorhombic to a tetragonal structure occurs via an intermediate dimer state as an orthorhombic phase to a dimer and then from this dimer to a tetragonal phase. In addition, the second process is slower than the first.

    When treated at higher P (500–800 °C at 5 GPa), C60 polymerized via [2 + 2] cycloaddition in the (111) plane of the original fcc phase into a rhombohedral structure with a hexagonal lattice.2,32 It has been reported that during decompression from 6 GPa to 2.2 GPa at 873 K, the rhombohedral polymer could transform to a tetragonal phase.40 A trigonal symmetry instead of a hexagonal one was confirmed to be the most stable mode by an XRD experiment.75 The hexagonal layers stack in a close-packed arrangement of the ABCABC type with space group R3-m.33,76 All the former four-membered rings rotate 60°, forming another structure with ACBACB stacking type. Single-crystal XRD studies confirm this structure,42 and a density functional theory (DFT) calculation demonstrates slightly more stability.77

    F. 3D polymerization of C60 at HPHT

    Under much higher P (>13 GPa) at elevated T, 3D polymers are obtained. For example, one superhard phase of a 3D polymer quenched from 13 GPa and 820 K features [3 + 3] cycloaddition bonding between the [2 + 2] cycloaddition-bonded 2D tetragonal layers.47,48 Single-crystal XRD results demonstrate that the 3D C60 polymer is in a body-centered orthorhombic space group Immm structure (Fig. 6).78 The resulting sp3–sp2 hybridized system is expected to show metallic conductivity. In contrast, the 2D rhombohedral phase transforms into a disordered phase at around 15 GPa,79 and the 2D tetragonal phase transforms into a 3D metastable polymer over 20 GPa at room temperature.80In situ Raman spectra strongly support the above conclusion by revealing a transformation from a 2D to a 3D phase.81–84 Also, the in situ X-ray measurements reveal a first-order irreversible transition of the tetragonal polymeric phase at around 24 GPa.82

    Crystal structure of the 3D C60 polymer (b) in comparison with the starting 2D polymer (a). Carbon atoms marked with ∗ in (c) are from the neighboring C60 units. Reprinted with permission from S. Yamanaka et al., Phys. Rev. Lett. 96, 076602 (2006). Copyright 2006 American Physical Society.

    Figure 6.Crystal structure of the 3D C60 polymer (b) in comparison with the starting 2D polymer (a). Carbon atoms marked with ∗ in (c) are from the neighboring C60 units. Reprinted with permission from S. Yamanaka et al., Phys. Rev. Lett. 96, 076602 (2006). Copyright 2006 American Physical Society.

    Although several studies have demonstrated the existence of 3D C60 polymers, there is still no consensus on the preparation conditions, owing to experimental difficulties and the structural complexity of the products. Generally, under 8–9 GPa, a cubic structure is observed up to about 600–700 K. Disorder increases, and covalent linking extends in all directions with rising T.5,68 Over 9.5 GPa, a body-centered structure presents.78 When T rises above 900–1000 K, the C60 cages start to collapse, and this, combined with the covalent bonding distortion of the cage, makes the structure become amorphous or even graphitic at higher T.4 Superhard materials, even cubic diamond crystallites, have also been found.85

    In conclusion, several tentative phase diagrams of C60 at HPHT have been reported.2,4,55,76 It should be noted that the polymeric phase formation and purity depend strongly on T, P, and treatment time. Typical conditions for the formation of C60 polymers are as follows: for the 1D orthorhombic phase, 1–1.2 GPa, 550–585 K for 5 h;39,55 for the 2D tetragonal phase, 2–2.5 GPa, 700–873 K for 0.5–4 h;42,43,55 and for the 2D rhombohedral phase, 5–6 GPa, 773–873 K for 0.5–1 h.44,55 Another problem is establishing the stability of the polymeric phases. A quenched intermediate transition or experimental artifacts all possibly exist, but only a few in situ studies have been carried out.

    It is important to highlight that heating-then-pressing from rapidly rotating monomers leads to disordered states, because chains form in all directions. Compression benefits the formation of well-ordered 1D or 2D polymers. Nevertheless, pressing-then-heating favors the optimal [2 + 2] cycloaddition route to dimerization, but orientational tuning at high T promotes linear or planar bonding.68 However, different paths still give distinct products. For instance, the heating-then-pressing path tends to form rhombohedral polymers.82

    When we consider the polymeric phase transition shown in Fig. 1, the orthorhombic-to-tetragonal transformation involves proper rotations of the chains followed by interchain cross-linking in the {100}c planes. Furthermore, the transformation from an orthorhombic to a rhombohedral phase demands that the chains rotate to a fixed angle, while the tetragonal-to-rhombohedral transition requires bond rearrangement (Fig. 5).73

    G. Properties of polymeric C60 phases

    The physical properties of the low-dimensional phases have already been well studied.2,4 The essential difference between the C60 monomer and the polymers is the covalent intermolecular linkage in the latter. 13C nuclear magnetic resonance (NMR) studies of orthorhombic,86,87 tetragonal,88 and rhombohedral86,89 C60 clearly demonstrate that these structural differences are based on the distinguishable frequencies of sp2 and sp3 hybridized carbon. The relative numbers of sp2 and sp3 hybridized atoms can be used to identify the intra- and intermolecular bonds in the polymeric phase. In particular, the sp2 peak can be decomposed into lines corresponding to the number of unique carbon atom environments (Fig. 7).88,89

    13C magic angle spinning (MAS) NMR spectra of the rhombohedral 2D phase with a spinning rate of 12 kHz and repetition times of 200 and 900 s. The inset shows the structural mode of this phase. Reprinted with permission from F. Rachdi et al., J. Phys. Chem. Solids 58, 1645–1647 (1997). Copyright 1997 Elsevier.

    Figure 7.13C magic angle spinning (MAS) NMR spectra of the rhombohedral 2D phase with a spinning rate of 12 kHz and repetition times of 200 and 900 s. The inset shows the structural mode of this phase. Reprinted with permission from F. Rachdi et al., J. Phys. Chem. Solids 58, 1645–1647 (1997). Copyright 1997 Elsevier.

    The reduced symmetry of icosahedral C60 and the formation of intermolecular bonds in the polymeric phase lead to splitting of degenerate molecular orbitals and a narrowing of the HOMO–LUMO gap. The electronic structures of 1D and 2D C60 polymers have been studied theoretically. The 1D linear polymer is a semiconductor with a finite band gap of 1.148 eV and almost pure s-type intermolecular bonding.90 The 2D tetragonal and rhombohedral C60 polymers are also semiconductors. With different intermolecular covalent bonds, the tetragonal polymeric and rhombohedral polymeric phases exhibit band gap values of 1.2 eV and 1.0 eV, respectively.34

    C60 is reportedly stable up to 950 °C under low-pressure inert gas conditions.91 High-temperature synchrotron radiation XRD results show that C60 is stable below temperatures of 900 °C at 80 MPa.92 However, all the polymers are less stable than the monomer with free C60 molecules. For example, dimers are stable to about 420 K,93,94 the orthorhombic phase breaks down near 500 K, and the tetragonal and rhombohedral 2D polymers are stable to 550 K.94,95 Thus, the stability sequence is dimer < 1D polymer < 2D polymer < monomer.34 It should be noted that the thermal “stability limit” is dependent on time scale. The activation energy Ea reflects the stability of the polymers; it is 1.75 eV and 1.9 eV for the dimers and orthorhombic C60 polymeric phases, respectively.34,94

    The degree of polymerization, the polymeric structure, and the ratio of sp2 to sp3 C dominate the conductivity of polymerized C60. In contrast to C60, an n-type semiconductor, the 1D plasma-polymerized C60 polymer acts as a p-type semiconductor in field effect transistor (FET) experiments owing to variable-range hopping (VRH).96 With the increasing degree of polymerization, the band gap reduces and the electrical conductivity increases.95 The conductivity of 2D C60 polymers is anisotropic. In the covalent bond linking layer, it shows a metallic-like behavior. In the weak van der Waals interacting direction of the interlayers, it exhibits semiconductor-like behavior.97,98 This phenomenon is ascribed to the weak localization of the carriers, and the metallic behavior could be attributed to a self-doping effect. Structural calculations show that lattice defects may induce metallic behavior.99 Electrical resistivity measurements show that amorphous polymeric C60 phases are semimetallic, with VRH and semiconducting behavior.100 Recent first-principles DFT calculations have demonstrated that the structures of 3D C60 polymers in which each molecule is in one of the two standard orientations are analogous to ordered binary-alloy-type structures.101 Furthermore, all the investigated structures in this case are predicted to be metallic, which is supported by an experimental result where a polymeric 56/56 2 + 2 cycloaddition bond is analogous to the Ising antiferromagnetic interaction in a 3D C60 polymer.102

    Thermal conductivity theoretically increases with increasing P. In addition, it increases with the formation of covalent bonds between near neighboring C60’s because the resulting appearance of new phonon modes contributes to heat transport. Polymeric phase disorder will compensate for the thermal conductivity effect.103

    III. HPHT RESEARCH ON C70

    C70 molecules undergo quasi-free rotation in a close-packed fcc lattice above room temperature. However, only the five bonds radiating out from each polar pentagon have significant reactivity, which leads to topological constraints on the formation of C70 polymers. The anisotropic C70 molecule exhibits at least three phases: free rotation at high T, uniaxial rotation at intermediate T, and frozen rotation or ratcheting at low T.3,104

    Under compression, the rotational motion is reduced to uniaxial rotation around the molecular axis. Subsequently, the C70 molecules tend to arrange in a line, and a rhombohedral structure is obtained (Fig. 8).106,107 Regarding symmetry, the C70 monomer with an fcc lattice cannot form a long-range ordered polymeric structure. In contrast, C70 crystallized in the hcp structure can polymerize into an ordered zig-zag chain structure with an orthorhombic lattice.108

    Structural phase diagram of C70 with an fcc structure under ambient conditions. Reprinted with permission from B. Sundqvist, Carbon 125, 258–268 (2017). Copyright 2017 LW Elsevier.

    Figure 8.Structural phase diagram of C70 with an fcc structure under ambient conditions. Reprinted with permission from B. Sundqvist, Carbon 125, 258–268 (2017). Copyright 2017 LW Elsevier.

    At T between 400 and 500 K and P around 1 GPa, a dimer-rich phase is observed, while the dimer ratio reduces at higher temperatures. Notably, C70 exhibits lower reactivity than C60,105 so the dimerization of C70 is much slower under high P.

    Several polymeric phases have been observed in C70 at higher P and T. Incompatibility between the lattice symmetry and the geometry of the reactive bonds means that an ordered polymeric phase of C70 is hard to obtain at P in the range of 0.82 GPa at T between 350 and 600 K. However, Blank et al.109,110 and Marques et al.111,112 reported that C70 with an fcc lattice could transform to an ordered structure under sufficiently high P, which was thought to occur only in hcp C70. It should be noted that the low reactivity and topochemical constraints on the polymerization lead to the slower formation of C70 oligomers and polymers. Above 700 K, a “disordered cross-linked layered” structure and a 3D polymerized tetragonal structure have been reported.109

    IV. HPHT RESEARCH ON INTERCALATED FULLERENE

    A. Solvent effect on parent structure at ambient conditions

    The interaction between fullerenes and solvents leads to the formation of solvated fullerene.113 C60 and C70 can form solvated crystals with almost all common solvents.12 Evaporating fullerene solutions is the general preparation method to obtain solvated fullerenes. In 1991, Fleming et al.114 reported solvated C60 and C70 for the first time, inspiring subsequent research on the synthesis and characterization of this series of materials.115–122 Several conclusions have been drawn based on this research:

    Solvates are typical van der Waals complexes with negative excess volumes. Bonding between fullerenes and a solvent will enhance the stability of the solvate.120

    Synthesis T plays an important role in the stability and stoichiometry of solvated fullerenes.14,15,120,123

    Incorporated solvent molecules influence the crystal structure. There are at least five types of packing symmetry (hexagonal in C60∗2BrCCl3,116 cubic in C60∗cyclohexane,118 monoclinic in C60∗C6H5CH3,123,124 triclinic in C60∗2ferrocene,122 and orthorhombic in C60∗1,1,2-trichloroethane115,125) in solvated fullerenes. An interesting case is a static cubane, which occupies the octahedral voids of the fcc structures and forms C60∗C8H8 and maintains the fcc structure (Fig. 9).49 Accordingly, solvents with symmetrical geometry and a small size tend to occupy the octahedral voids of fullerene lattices and maintain their initial structure, while solvents with non-symmetric geometry or polar molecules are ideal structure controllers.12,120,123,124,126 Besides the solvent species, the stoichiometric ratio of solvent to C60 will also have an impact on the structure of the solvated crystals.116

    Structures of fullerene–cubane heteromolecular crystals. (a) Space-filling view of the rotor–stator phase of C60∗C8H8. Rotating fullerenes occupy the fcc lattice sites, and static cubanes are in between the octahedral voids. (b) The position of cubane in an octahedral void of the fcc cell. The concave outer faces show the perfect match between the cubane and the surrounding fullerenes. (c) The restricted motion of C70 in the tetragonal rotor–stator phase of C70∗C8H8. In the molecular bearing of six cubanes, C70 rotates about its long axis (C5), which processes about the c axis of the body-centered tetragonal cell. Reprinted with permission from S. Pekker et al., Nat. Mater. 4, 764–767 (2005). Copyright 2005 Springer Nature.

    Figure 9.Structures of fullerene–cubane heteromolecular crystals. (a) Space-filling view of the rotor–stator phase of C60∗C8H8. Rotating fullerenes occupy the fcc lattice sites, and static cubanes are in between the octahedral voids. (b) The position of cubane in an octahedral void of the fcc cell. The concave outer faces show the perfect match between the cubane and the surrounding fullerenes. (c) The restricted motion of C70 in the tetragonal rotor–stator phase of C70∗C8H8. In the molecular bearing of six cubanes, C70 rotates about its long axis (C5), which processes about the c axis of the body-centered tetragonal cell. Reprinted with permission from S. Pekker et al., Nat. Mater. 4, 764–767 (2005). Copyright 2005 Springer Nature.

    Furthermore, the solvent molecule also affects the physical properties. Świetlik et al.123 and Graja and Świetlik127 reported that the rotation of C60 is hindered by the intercalated solvent molecule, which is reflected in the vibrational spectrum. The influence on the rotation of C60 depends dramatically on the intermolecular interaction and structural arrangement. On the other hand, the symmetry decrease due to interactions between fullerene and solvent molecules enhances the photoluminescent intensity. For instance, the intensity rises about two orders of magnitude in C60m-xylene compared with pristine C60 crystal,12,14 and it is about 30 times higher in C70∗mesitylene than in pristine C70 powder.128 The solvent effect on the photoluminescent behavior correlates with the interaction of the cages.129

    B. Solvent effect under HPHT

    Under high P, the incorporated solvent molecule diversified the phase transformation and polymerization of C60. There are four kinds of P-induced solvent effect:

    Through a spacer, such as C8H8 in the C60∗C8H8 crystal, which effectively shields the P in the interball voids, raising the phase transition P compared with that of C60.130

    Through a structure confiner, for instance, 1,1,2-trichloroethane (TCAN), which confines the [2 + 2] cycloaddition bonds of the C60’s that form in the upper and lower layers alternately when C60∗1TCAN is treated under high P and high T. As a result, a novel quasi-3D C60 polymer is obtained.17 Another example is C60m-xylene.16,131 The long-range hcp structure is maintained by the m-xylene confiner even when the C60 cages collapse at P > 32 GPa.

    Through a bridge, such as in C70m-xylene, where the m-xylene forms a superhard sp3-bonded carbon by bonding with neighboring fullerenes.83

    Through a charge reservoir. For example, the formation of a donor–acceptor complex can be monitored by Raman spectroscopy when C60{Ni(nPr2dtc)2}, C60{Cu(nPr2dtc)2}, and C60{Pt(dbdtc)2} are compressed at P > 0.7 GPa, 0.7 GPa, and 0.5 GPa, respectively.132,133 In the case of iodine-doped C60, which crystallizes in a base-centered orthorhombic structure,134P enhances the charge transfer between the C60 and the iodine atoms. As a result, the electrical band gap is 0.3 eV at 17 GPa.

    It must also be mentioned that any given solvent molecule generally plays more than one role in the structure control and property tuning of solvated fullerene. For instance, ferrocene acts as a spacer as well as a charge reservoir to tune the polymerization process and promotes a layered structure of the intercalated ferrocene in the C70(Fe(C5H5)2)2 crystal.135 The m-xylene in a C60 solvate plays an important role as a spacer and a bridge by forming covalent bonds with neighbors. It preserves the stability of the highly deformed or amorphous C60 molecules to form a long-range ordered structure under high P.83 Hence, the solvent effects on the structure and properties of solvated molecules are various and diverse under high P.

    The P-induced polymeric products can be divided into two categories: copolymerized products of C60 and solvent molecules, or C60 self-condensed materials. C60∗C8H8 and C60S16 are in the former category. An anomaly at 0.5 GPa in the P dependences of the molecular vibrational frequencies indicates an orientational ordering transition at room temperature, while an anomaly at 1.3 GPa is attributed to a fullerene–cubane interaction in C60∗C8H8.136 In C60S16, the peaks from the S8 rings disappear, and new peaks are observed above 7.2 GPa, suggesting that the sulfur rings break and covalent C–S bonds form between the C60 and solvent molecules.137 In the second category, the C60 solvates grown from different solvent molecules exhibit various polymeric products under high P. For instance, C60∗TDAE [where TDAE is tetrakis(dimethylamino)ethylene] (Refs. 138 and 139) and C60(Fe(C5H5)2)2 (Refs. 140 and 141) both exhibit a one-dimensional polymerized phase above ∼1 GPa and 5 GPa, respectively. C60{Pt(dbdtc)2} starts to form intercage covalent bonds at 0.5 GPa and a novel polymer, which is coordinated with nearest-neighbor C60 with 10 and 6 sp3-like carbon atoms per molecule above 2.5 GPa.132 The solvent molecule in C60m-xylene confines the formation of ordered amorphous carbon clusters (OACCs) under high P, as shown in Fig. 10.16 This study on a small piece of C60 solvate indicates the diversity of the structure and properties of these solvates under high P, and additionally predicts potential applications in a wide range of areas.

    Simulated structures of C60∗m-xylene under different compression and decompression conditions. Far left, the pristine structure at 0.4 GPa. The deformation of the C60 cages is elastic below about 30 GPa, and the deformed cages return to their initial shape as they are decompressed back to ambient P. Above 30 GPa, many carbon–carbon bonds start to break, and the cages can no longer return to C60 upon decompression. OACCs retain long-range periodicity and can be preserved under ambient P conditions. Reprinted with permission from L. Wang et al., Science 337, 825–828 (2012). Copyright 2012 The American Association for the Advancement of Science.

    Figure 10.Simulated structures of C60m-xylene under different compression and decompression conditions. Far left, the pristine structure at 0.4 GPa. The deformation of the C60 cages is elastic below about 30 GPa, and the deformed cages return to their initial shape as they are decompressed back to ambient P. Above 30 GPa, many carbon–carbon bonds start to break, and the cages can no longer return to C60 upon decompression. OACCs retain long-range periodicity and can be preserved under ambient P conditions. Reprinted with permission from L. Wang et al., Science 337, 825–828 (2012). Copyright 2012 The American Association for the Advancement of Science.

    Recently, a new way to understand the pressure-induced structure formation and transformation in fullerene solvates under high P has been proposed by studies in which their building blocks have been tailored and their boundary interactions have been tuned.83,84,142,143 Solvent molecules with hexagonal carbon rings, such as m-xylene, 1,2,4-trimethylbenzene (TMB), and m-dichlorobenzene (m-Cl), tend to stabilize highly compressed or collapsed fullerene clusters in fullerene solvates and form a long-range ordered structure under compression. In contrast, ferrocene, with its pentagonal carbon ring, generally causes a transformation into an amorphous component together with a carbon cage and leads to the formation of a disordered fullerene solvate structure under high P. Cubane, C8H8, differs from these aromatic solvents. It reacts with C60 in solvates under high P, but not all of the C8H8 molecules react with the carbon cage.142 The intercalated C8H8 and the reacted C8H8 both protect the C60 cages from early collapse under compression. Furthermore, unlike C60m-xylene, in which the C60 molecules overcome the confining effect of the solvent molecules and take part in polymerization between the cages,16,84 the absence of intermolecular bonds between the C60 molecules in C8H8/C60 tunes the boundary interactions of the carbon clusters to retain an ordered arrangement up to 45 GPa.

    C. HPHT studies on CNT-confined fullerenes

    Confinement of fullerene molecules by the intrinsic structure of a carbon nanotube (CNT) in HPHT synthesis is an important strategy for the polymerization of fullerenes.144–150 C60’s packed inside single-walled CNTs are considered “peapods”145 and provide an opportunity to study original quasi-1D fullerene phases. Using high-resolution transmission electron microscopy, Smith and co-workers145,146 discovered self-assembling C60 fullerene chains inside CNTs. Molecular dynamics studies suggest that (9,9) and (10,10) nanotubes with diameter larger than 6.27 Å give the best fit for C60.147,148 The intercage distance in a CNT was reduced to 8.7 Å under compression to 25 GPa, and a one-dimensional polymer chain was obtained.20,52 Larger CNTs have more space for the dimensional polymerization of C60, as proven by the observed zigzag structure in C60@CNT(15,15).

    In contrast, the anisotropy of C70 is reflected by two distinct orientations with regard to the nanotube axis. The C70 long axis parallel to the nanotube long axis leads to a lying-down orientation, while perpendicular orientations form when the long axis is rotated 90° to the nanotube axis. The orientation of C70 depends on the diameter of the nanotube. C60 peapods fully polymerize at 1.5 GPa and 300 °C,149 while C70 peapods do not give a polymeric product even at higher P and T (1.5 GPa and 545 °C, or 2.5 GPa and 300 °C).104 These results show that CNTs impose much stricter constraints on the reactivity of the ovoid C70 than on that of the high-symmetry C60.

    Theoretical simulations have predicted several structures of fullerene peapods under high P.151 In 2017, a novel superhard sp3 carbon allotrope was obtained by cold compression of C70 peapods.150 This so-called V carbon is thermodynamically stable over a wide P range up to 100 GPa and can be recovered to ambient conditions once it is formed, providing a new strategy for synthesizing new superhard materials.

    D. HPHT studies on metal-doped fullerenes

    Polymerized alkali-metal-doped fullerenes have attracted much attention because of their hardness and stiffness, metallic behavior,152,153 and especially superconductivity.154 1D and 2D polymeric alkali-metal-doped fullerenes can be obtained under HPHT. For example, AC60 (A = K, Rb, Cs) polymerized to a 1D orthorhombic phase under high T.155 At 5 GPa and 573 K, the 2D rhombohedral polymers (LiC60)HP and (NaC60)HP were synthesized starting from C60 and AN3 (A = Li, Na). Notably, large-radius alkali metals (Rb or Cs) and stoichiometric X6C60 stabilized the doped fullerene at much higher P compared with pristine C60. Cs6C60 retains a monomeric structure until 45 GPa,156 while Rb6C60 exhibits a reversible structural transition to a 2D polymeric phase at 35 GPa and 600 K.157 A 3D polymeric fulleride can be synthesized by compressing 2D Na4C60.21 On the other hand, fullerene will polymerize in a lower P range when doped with smaller alkali metals such as Na or Li.153,154 In addition, an intercalated alkali metal with a small radius could diffuse in the fullerene lattice such as in the case of the superionic conductivity in the Li4C60 polymer,158,159 Na-ion diffusion in Na2C60 results in the formation of polymeric phases of C60.160

    V. HPHT RESEARCH ON ENDOHEDRAL FULLERENES

    As pointed out at the start of this paper, pressure is a fundamental thermodynamic variable that can dramatically alter the atomic and electronic structure of materials, resulting in concomitant changes in their properties. Determining the evolution of the geometric and electronic structure of endohedral fullerenes under high P is particularly important because this information is necessary for understanding the electronic conduction mechanism, for discoveries of novel electronic properties, and for fabricating novel fulleride-based materials. On the other hand, the relatively low yield restricts the characterization of endohedral metallofullerenes by standard measurement techniques, while our limited understanding of cluster structure and properties is another challenge. However, in situ measurements at high P in a diamond anvil cell offer a complete workflow to circumvent these problems effectively.

    Cui et al.23,24 have conducted high-P structural transformation studies on the mono-atom encapsulated fullerenes Sm@C88 and solvated Sm@C90. The high-P in situ IR spectra exhibit both red and blue shifts, indicating an anisotropic deformation of the carbon cage under compression. On the basis of these results, together with those of simulations, Cui et al. propose that the cage deforms in the vertical direction, changing from an ellipsoidal to an approximately spherical shape. With further compression, a peanut-like structure appears, before the cage finally collapses under higher P. At the same time, there is a significant P-induced reduction in the band gap because of the deformation of the carbon cage and the enhanced intermolecular interaction (Fig. 11). In this compression process, the Sm atom plays a role in restraining the compression of adjacent bonds. The structural evolution of solvated Sm@C90 under high P is almost the same.24 The solvent molecule plays a role protecting against cage collapse and postpones P-induced changes, contributing to the trapping of the metal atom in an OACC.

    (a) The mid-IR region shows the absorption edge under high P. (b) The band gap as a function of P. The inset shows plots of (αhν)2 versus hν at ambient P and at 13.9 GPa. (c) IR reflectivity spectra of the sample under high P. Reprinted with permission from J. Cui et al., Sci. Rep. 5, 13398 (2015). Copyright 2015 Springer Nature.

    Figure 11.(a) The mid-IR region shows the absorption edge under high P. (b) The band gap as a function of P. The inset shows plots of (αhν)2 versus at ambient P and at 13.9 GPa. (c) IR reflectivity spectra of the sample under high P. Reprinted with permission from J. Cui et al., Sci. Rep. 5, 13398 (2015). Copyright 2015 Springer Nature.

    Polymeric fullerenes such as polymerized C60 and C70 exhibit unique properties such as high elastic constants and hardness.161 An HPHT-induced polymerization study of the endohedral fullerene La@C82 revealed different polymeric phases such as dimers and 2D and 3D polymers.22 X-ray and electron diffraction results showed increasing density and disorder when the sample was treated at 9.5 GPa and 520 K. A fraction of dimer and oligomers mixed with mainly monomeric La@C82 retained the hcp structure of the pristine sample. The hcp lattice constant decreased dramatically, and small amounts of body-centered tetragonal phase appeared when T was raised to 720 K.

    VI. CONCLUSION

    HPHT studies on fullerene and its derivatives have been reviewed. The essential differences among the polymeric phases concern the orientational phenomena that they exhibit and the fullerene cage bonding modes. In view of these, various 1D, 2D, and 3D C60 polymeric structures prepared under HPHT have been studied, utilizing FTIR and Raman spectroscopy and XRD. C70, a nonspherical carbon cage with lower symmetry than C60, shares the general features of the latter with regard to polymerization, but its polymeric structures exhibit significantly fewer phases. A deeper understanding of the polymerization processes and phase diagrams of the fullerenes will help promote the synthesis of new polymeric forms with interesting and useful structural, electrical, magnetic, and photoluminescence properties.

    Based on a molecular engineering strategy, solvent molecules, CNTs, and metal atoms could be used to control the arrangement of carbon cages in the lattice. Under HPHT, the intercalated units dominate the evolution of the fullerene configuration, constrain polymerization behavior, and determine the structure and properties of the polymeric product. Further efforts are required so that the phase diagrams of these intercalated/doped fullerenes can be drawn. In addition, studies of intercalated/doped fullerenes at high P and very high T (>1000 °C) present exciting opportunities to discover unique carbon phases and even diamond-like polymeric structures.

    Endohedral atoms/clusters greatly enrich the properties of fullerenes, and the resulting fullerene derivatives have properties ranging from semiconducting to metallic. In particular, the tunable electronic spin state inside the fullerene cage makes endohedral fullerenes candidates for use in quantum computing. The application of high P is a potentially useful method to control electronic spin and phonon behavior. However, to date, there have been few high-P studies of fullerenes that have considered this topic. Study of endohedral fullerenes at high P is a very recent field and much remains unexplored. What happens to the interplay of carbon cages and the endohedral species under HPHT conditions? How is charge transferred between the carbon cages and the encaged species when the density increases under compression? What unique properties due to the evolution of molecular and electronic structure will endohedral fullerenes exhibit under high P and indeed high T? The need to answer these questions is a strong motivation for further research in this field.

    References

    [1] L. Wang, H. Mao, Y. Ding, X. Chen, B. Li. Solids, liquids, and gases under high pressure. Rev. Mod. Phys., 90, 015007(2018).

    [2] B. Sundqvist. Fullerenes under high pressures. Adv. Phys., 48, 1-134(1999).

    [3] B. Sundqvist. Polymeric fullerene phases formed under pressure. Struct. Bonding, 109, 85-126(2004).

    [4] G. A. Dubitsky, N. R. Serebryanaya, S. G. Buga, P. M. Yu, V. D. Blank, B. Sundqvist. High-pressure polymerized phases of C60. Carbon, 36, 319-343(1998).

    [5] R. A. Wood, M. H. Lewis, N. Kitamura, S. M. Bennington, A. K. Fukumi, M. G. Cain et al. In situ diffraction measurement of the polymerization of C60 at high temperatures and pressures. J. Phys.: Condens. Matter, 12, L451-L456(2000).

    [6] A. Onodera, M. Kobayashi, K. Suito, T. Horikawa. Time-resolved x-ray diffraction study of C60 at high pressure and temperature. Phys. Lett. A, 287, 143-151(2001).

    [7] T. Le Bihan, L. S. Dubrovinsky, A. V. Talyzin, U. Jansson. Pressure-induced polymerization of C60 at high temperatures: An in situ Raman study. Phys. Rev. B, 65, 245413(2002).

    [8] Y. Hou, B. Liu, D. Liu, B. Zou, S. Yu, L. Wang et al. High pressure and high temperature induced polymeric C60 nanocrystal. Diamond Relat. Mater., 17, 620-623(2008).

    [9] B. Liu, M. Yao, L. Wang, S. Yu, Y. Hou, A. Chen et al. Comparative study of pressure-induced polymerization in C60 nanorods and single crystals. J. Phys.: Condens. Mater., 19, 425207(2007).

    [10] S. Yu, L. Wang, A. Chen, M. Yao, B. Liu, Y. Hou et al. Photoluminescence properties of high-pressure-polymerized C60 nanorods in the orthorhombic and tetragonal phases. Appl. Phys. Lett., 89, 181925(2006).

    [11] Y. Hou, Q. Zhao, B. Liu, T. Cui, L. Wang, H. Ma et al. Pressure-induced polymerization of nano- and submicrometer C60 rods into a rhombohedral phase. Chem. Phys. Lett., 423, 215-219(2006).

    [12] L. Wang. Solvated fullerenes, a new class of carbon materials suitable for high-pressure studies: A review. J. Phys. Chem. Solids, 84, 85-95(2015).

    [13] L. Wang, M. Yao, Y. Hou, B. Liu, S. Yu, D. Liu et al. Synthesis and high pressure induced amorphization of C60 nanosheets. Appl. Phys. Lett., 91, 103112(2007).

    [14] B. Liu, D. Liu, L. Wang, S. Yu, M. Yao, Y. Hou et al. Highly enhanced luminescence from single-crystalline C60•1m-xylene nanorods. Chem. Mater., 18, 4190-4194(2006).

    [15] M. Yao, S. Yu, Y. Hou, L. Wang, B. Liu, D. Liu et al. Synthesis of thin, rectangular C60 nanorods using m-xylene as a shape controller. Adv. Mater., 18, 1883-1888(2006).

    [16] W. Yang, H. Li, S. Sinogeikin, B. Liu, Y. Ding, L. Wang et al. Long-range ordered carbon clusters: A crystalline material with amorphous building blocks. Science, 337, 825-828(2012).

    [17] M. Yao, M. Feng, Y. Yuan, X. Li, C. Pei, Z. Yang et al. Quasi 3D polymerization in C60 bilayers in a fullerene solvate. Carbon, 124, 499-505(2017).

    [18] M. Yao, S. Sun, B. Sundqvist, W. Cui, B. Liu. High pressure and high temperature induced polymerization of doped C60 materials. Carbon, 109, 269-275(2016).

    [19] R. Liu, M. Yao, W. Cui, Q. Li, B. Zou, D. Liu et al. Pressure induced metastable polymerization in doped C60 materials. Carbon, 115, 740-745(2017).

    [20] R. Almairac, M. Chorro, L. Alvarez, J.-L. Sauvajol, S. Rols, J. Cambedouzou et al. Structural properties of carbon peapods under extreme conditions studied using in situ x-ray diffraction. Phys. Rev. B, 74, 205425(2006).

    [21] V. Pischedda, B. Sundqvist, R. Debord, T. Wågberg, M. Mezouar, M. Yao et al. Pressure-induced transformation in Na4C60 polymer: X-ray diffraction and Raman scattering experiments. Phys. Rev. B, 84, 144106(2011).

    [22] B. A. Kulnitskiy, N. R. Serebryanaya, G. Liu, S. G. Buga, M. Klaeser, V. D. Blank et al. Structure and properties of solid La@C82 endofullerene polymerized under pressure 9.5 GPa and temperature 520-720 K. Synth. Met., 121, 1093-1096(2001).

    [23] H. Yang, J. Cui, F. Ma, M. Yao, Q. Li, Z. Liu et al. Structural deformation of Sm@C88 under high pressure. Sci. Rep., 5, 13398(2015).

    [24] Z. Liu, H. Yang, M. Du, S. Liu, J. Cui, M. Yao et al. Structural stability and deformation of solvated Sm@C2(45)-C90 under high pressure. Sci. Rep., 6, 31213(2016).

    [25] Z. Yao, Z. Yan, C. Pei, S. Samanta, H. Xiao, G. Liu et al. Pressure-induced anisotropy deformation of fullerene cages and pyramidalization of planar endohedral clusters.

    [26] N. Martín, F. Giacalone. Fullerene polymers: Synthesis and properties. Chem. Rev., 106, 5136-5190(2006).

    [27] J. M. Holden, Y. Wang, K. Wang, A. M. Rao, P. Zhou, G. T. Hager et al. Photoinduced polymerization of solid C60 films. Science, 259, 955-957(1993).

    [28] K. Suzuki, S. Yamamoto, A. Hayashi, T. Matsuoka. The first application of fullerene polymer-like materials, C60Pdn, as gas adsorbents. J. Mater. Chem., 14, 2633-2637(2004).

    [29] L. Marques, J.-L. Hodeau, M. Núñez-Regueiro, A. M. Rao, P. C. Eklund. Infrared and Raman studies of pressure-polymerized C60. Phys. Rev. B, 55, 4766-4773(1997).

    [30] N. Takahashi, N. Matsuzawa, M. Ata, H. Dock. Plasma-polymerized C60/C70 mixture films: Electric conductivity and structure. J. Appl. Phys., 74, 5790(1993).

    [31] X. W. Zhang, B. Wang, Y. L. Li, H. Yan, J. Z. Cui, Y. J. Zou et al. Bonding character of the boron-doped C60 films prepared by radio frequency plasma assisted vapor deposition. J. Mater. Sci., 37, 1043-1047(2002).

    [32] T. Siegrist, Y. Iwasa, R. M. Fleming, R. C. Haddon, O. Zhou, T. Arima et al. New phases of C60 synthesized at high pressure. Science, 264, 5170-5172(1994).

    [33] O. Béthoux, J.-L. Hodeau, M. Núñez-Regueiro, L. Marques, M. Perroux. Polymerized fullerite structures. Phys. Rev. Lett., 74, 278-281(1995).

    [34] C. Xu, G. E. Scuseria. Theoretical predictions for a two-dimensional rhombohedral phase of solid C60. Phys. Rev. Lett., 74, 274-277(1995).

    [35] T. Wågberg, R. Moret, B. Sundqvist. Influence of the pressure-temperature treatment on the polymerization of C60 single crystals at 2 GPa-700 K. Carbon, 43, 709-716(2005).

    [36] U. Edlund, P. Jacobsson, D. Johnels, B. Sundqvist, P. Launois, J. Jun et al. Structural and physical properties of pressure polymerized C60. Carbon, 36, 657-660(1998).

    [37] R. K. Nikolaev, V. D. Negrii, A. P. Moravsky, A. N. Izotov, I. O. Bashkin, Yu. A. Ossipyan et al. Photoluminescence of solid C60 polymerized under high pressure. Chem. Phys. Lett., 272, 32-37(1997).

    [38] A. N. Andriotis, R. M. Sheetz, L. Chernozatonskii, M. Menon. Magnetic properties of C60 polymers. Phys. Rev. Lett., 90, 026801(2003).

    [39] P. Launois, B. Sundqvist, P.-A. Persson, R. Moret. First x-ray diffraction analysis of pressure polymerized C60 single crystals. Europhys. Lett., 40, 55-60(1997).

    [40] V. A. Davydov, L. S. Kashevarova, V. Agafonov, H. Allouchi, A. V. Rakhmanina, R. Céolin et al. Tetragonal polymerized phase of C60. Phys. Rev. B, 58, 14786-14790(1998).

    [41] V. Agafonov, R. Céolin, H. Allouchi, A. V. Dzyabchenko, V. A. Davydov, H. Szwarc. Tetragonal polymerized phase of C60: Experimental artifact or reality?. Synth. Met., 103, 2415-2416(1999).

    [42] B. Sundqvist, R. Moret, T. Wågberg, P. Launois. High-pressure synthesis, structural and Raman studies of a two-dimensional polymer crystal of C60. Eur. Phys. J. B, 15, 253-263(2000).

    [43] X. Chen, S. Yamanaka. Single-crystal x-ray structural refinement of the ‘tetragonal’ C60 polymer. Chem. Phys. Lett., 360, 501-508(2002).

    [44] Y. Inoue, X. Chen, S. Yamanaka, M. Yasukawa, K. Sako. First single-crystal x-ray structural refinement of the rhombohedral C60 polymer. Chem. Phys. Lett., 356, 291-297(2002).

    [45] I. V. Trushkov, M. A. Yurovskaya. Cycloaddition to buckminsterfullerene C60: Advancements and future prospects. Russ. Chem. Bull., 51, 367-443(2002).

    [46] L. S. Kashevarova, N. Oleinikov, V. A. Davydov, V. M. Senyavin, A. V. Rakhmanina, O. P. Pronina et al. Pressure-induced dimerization kinetics of fullerene C60. J. Exp. Theor. Phys. Lett., 72, 557-560(2000).

    [47] N. R. Serebryanaya, L. A. Chernozatonskii, B. N. Mavrin. The superhard crystalline three-dimensional polymerized C60 phase. Chem. Phys. Lett., 316, 199-204(2000).

    [48] N. R. Serebryanaya, L. A. Chernozatonskii. Modelling and interpretation of the experimental data on the 3D polymerized C60 fullerites. Solid State Commun., 114, 537-541(2000).

    [49] G. Klupp, G. Bényei, S. Pekker, É. Kováts, G. Oszlányi, G. Bortel et al. Rotor-stator molecular crystals of fullerenes with cubane. Nat. Mater., 4, 764-767(2005).

    [50] J. L. Hodeau, M. Álvarez-Murga. Structural phase transitions of C60 under high-pressure and high-temperature. Carbon, 82, 381-407(2015).

    [51] V. N. Agafonov, A. V. Rakhmanina, A. V. Dzyabchenko, V. A. Davydov, P. Dubois, L. S. Kashevarova et al. Identification of the polymerized orthorhombic phase of C60 fullerene. J. Exp. Theor. Phys. Lett., 66, 120-125(1997).

    [52] T. Hara, F. Okino, H. Kataura, H. Touhara, T. Yokomae, S. Kawasaki et al. Pressure-polymerization of C60 molecules in a carbon nanotube. Chem. Phys. Lett., 418, 260-263(2006).

    [53] G. Oszlányi, G. Baumgartner, L. Forró, G. Faigel. Na4C60: An alkali intercalated two-dimensional polymer. Phys. Rev. Lett., 78, 4438-4441(1997).

    [54] C. R. Wang, K. Tan, E.-Y. Zhang, X. Lu, N. Chen. Size effect of encaged clusters on the exohedral chemistry of endohedral fullerenes:  A case study on the pyrrolidino reaction of ScxGd3-xN@C80 (x = 0 − 3). Org. Lett., 9, 2011-2013(2007).

    [55] H. Szwarc, V. M. Senyavin, L. S. Kashevarova, R. Céolin, A. V. Rakhmanina, V. A. Davydov et al. Spectroscopic study of pressure-polymerized phases of C60. Phys. Rev. B, 61, 11936(2000).

    [56] P. Zhou, M. S. Dresselhaus, P. C. Eklund, G. Dresselhaus, J. M. Holden, Z. Dong. Observation of higher-order Raman modes in C60 films. Phys. Rev. B, 48, 2862-2865(1993).

    [57] S. Ren, K. Wang, G. T. Hager, P. Zhou, J. M. Holden, Y. Wang et al. Raman scattering in C60 and alkali-metal-doped C60 films. Phys. Rev. B, 45, 1955-1958(1992).

    [58] K. Wang, P. Zhou, M. S. Dresselhaus, P. C. Eklund, Y. Wang, G. Dresselhaus et al. Raman scattering in C60 and alkali-metal-saturated C60. Phys. Rev. B, 46, 2595-2605(1992).

    [59] D. S. Bethune, G. Meijer, H. J. Rosen, W. C. Tang. The vibrational Raman spectra of purified solid films of C60 and C70. Chem. Phys. Lett., 174, 219-222(1990).

    [60] R. C. Haddon, A. F. Hebard, S. Glarum, S. J. Duclos. Raman studies of alkali-metal doped AxC60 films (A = Na, K, Rb, and Cs; x = 0, 3, and 6). Science, 254, 1625-1627(1991).

    [61] G. Dresselhaus, A. M. Rao, K.-A. Wang, M. S. Dresselhaus, P. C. Eklund. Observation of higher-order infrared modes in solid C60 films. Phys. Rev. B, 48, 11375-11380(1993).

    [62] T. Pichler, H. Kuzmany, R. Winkler. Infrared spectroscopy of fullerenes. J. Phys.: Condens. Matter, 7, 6601-6624(1995).

    [63] P. Jacobsson, B. Sundqvist, T. Wågberg. Comparative Raman study of photopolymerized and pressure-polymerized C60 films. Phys. Rev. B, 60, 4535-4538(1999).

    [64] T. Wågberg, P.-A. Persson, P. Jacobsson, B. Sundqvist. A Raman study of polymerised C60. Appl. Phys. A, 64, 223-226(1997).

    [65] Z.-T. Zhu, K. Kamarás, J. B. Page, G. B. Adams, J. L. Musfeldt, V. A. Davydov et al. Far-infrared vibrational properties of tetragonal C60 polymer. Phys. Rev. B, 65, 085413(2002).

    [66] G. Krexner, L. Pintschovius, O. Blaschko, N. Pyka. Bulk modulus of C60 studied by single-crystal neutron diffraction. Phys. Rev. B, 59, 11020-11026(1999).

    [67] T. Wågberg, R. Moret, B. Sundqvist, V. A. Davydov, V. Agafonov, P. Launois et al. Single-crystal structural study of the pressure-temperature-induced dimerization of C60. Eur. Phys. J. B, 37, 25-37(2004).

    [68] R. Moret. Structures, phase transitions and orientational properties of the C60 monomer and polymers. Acta Crystallogr., Sect. A: Found. Adv., 61, 62-76(2005).

    [69] A. V. Rakhmanina, V. Agafonov, V. A. Davydov, H. Allouchi, L. S. Kashevarova, R. Céolin et al. Particularities of C60 transformations at 1.5 GPa. J. Phys. Chem. B, 103, 1800-1804(1999).

    [70] K. R. Subbaswamy, M. Sawtarie, M. Menon. Structure and properties of C60 dimers by generalized tight-binding molecular dynamics. Phys. Rev. B, 49, 13966-13969(1994).

    [71] A. F. Gurov, I. O. Bashkin, N. P. Kobelev, A. P. Moravsky, O. G. Rybchenkot, V. I. Rashchupkin. A new phase transition in the T-P diagram of C60 fullerite. J. Phys.: Condens. Matter, 6, 7491(1994).

    [72] L. S. Kashevarova, A. V. Rakhmanina, A. Kahn-Harari, P. Dubois, V. A. Davydov, V. Agafonov et al. ‘Low-pressure’ orthorhombic phase formed from pressure-treated C60. Chem. Phys. Lett., 267, 193-198(1997).

    [73] B. Sundqvist, R. Moret, T. Wågberg, P. Launois. Chain orientation and layer stacking in the high-pressure polymers of C60: Single crystal studies. AIP Conf. Proc., 544, 81-84(2000).

    [74] A. Soldatov, B. Sundqvist, T. Wågberg. Spectroscopic study of phase transformations between orthorhombic and tetragonal C60 polymers. Eur. Phys. J. B, 49, 59-65(2006).

    [75] R. Céolin, A. V. Dzyabchenko, H. Szwarc, V. A. Davydov, V. Agafonov. Packing models for high-pressure polymeric phases of C60. J. Solid State Chem., 141, 164-167(1998).

    [76] J. G. Hou, T. Huang, S. Lu, A. D. Zhao. C60-based materials. ChemInform, 36(2005).

    [77] A. Oshiyama, S. Okada. Electronic structure of metallic rhombohedral C60 polymers. Phys. Rev. B, 68, 235402(2003).

    [78] A. Kubo, S. Yamanaka, K. Inumaru, K. Komaguchi, T. Inoue, N. S. Kini et al. Electron conductive three-dimensional polymer of cuboidal C60. Phys. Rev. Lett., 96, 076602(2006).

    [79] J. Arvanitidis, K. P. Meletov, G. A. Kourouklis, K. Prassides, Y. Iwasa. Pressure-induced transformation and phonon modes of the two-dimensional rhombohedral polymer of C60: A Raman spectroscopy. Phys. Rev. B, 68, 094103(2003).

    [80] J. Haines, V. Agafonov, V. A. Davydov, J. M. Léger. Irreversible amorphization of tetragonal two-dimensional polymeric C60 under high pressure. Solid State Commun., 121, 241-244(2002).

    [81] I. Tsilika, G. A. Kourouklis, S. Ves, S. Assimopoulos, J. Arvanitidis, K. P. Meletov et al. High-pressure-induced metastable phase in tetragonal 2D polymeric C60. Chem. Phys. Lett., 341, 435-441(2001).

    [82] L. S. Dubrovinsky, A. V. Talyzin. In situ Raman study of path-dependent C60 polymerization: Isothermal compression up to 32 GPa at 800 K. Phys. Rev. B, 68, 233207(2003).

    [83] W. Cui, Q. Li, F. Ma, R. Liu, S. Liu, M. Yao et al. A new carbon phase constructed by long-range ordered carbon clusters from compressing C70 solvates. Adv. Mater., 26, 7257-7263(2014).

    [84] M. Yao, J. Cui, W. Cui, R. Liu, S. Chen, J. Xiao et al. Pressure-induced transformation and superhard phase in fullerenes: The effect of solvent intercalation. Appl. Phys. Lett., 103, 071913(2013).

    [85] B. Bouchet-Fabre, J. M. Tonnerre, J. L. Hodeau, M. Perroux, J. J. Capponi, M. Núñez-Regueiro. High-pressure transformations of C60 to diamond and sp3 phases at room temperature and to sp2 phases at high temperature. Phys. Rev. B, 50, 10311-10314(1994).

    [86] M. Núñez-Regueiro, J.-L. Hodeau, F. Rachdi, L. Marques, C. Goze, L. Hajji et al. High resolution 13C NMR studies of one- and two-dimensional polymerized C60 under high pressure. J. Phys. Chem. Solids, 58, 1645-1647(1997).

    [87] P. Jacobsson, D. Johnels, B. Sundqvist, A. Soldatov, P.-A. Persson, U. Edlund. NMR and Raman characterization of pressure polymerized C60. Chem. Phys. Lett., 258, 540-546(1996).

    [88] V. A. Davydov, V. Agafonov, Y. Errammach, A. Rezzouk, F. Rachdi. High-resolution 13C NMR studies of the tetragonal two-dimensional polymerized C60 phase. Physica E, 8, 1-4(2000).

    [89] J.-L. Hodeau, L. Hajji, L. Marques, M. Núñez-Regueiro, C. Goze, F. Rachdi et al. High-resolution 13C NMR studies of high-pressure-polymerized C60: Evidence for the [2+2] cycloaddition structure in the rhombohedral two-dimensional C60 polymer. Phys. Rev. B, 54, R3676-R3678(1996).

    [90] T. Yamabe, Y. Oshima, K. Tanaka, Y. Matsuura. Electronic structure of a linear C60 polymer. Solid State Commun., 93, 163-165(1995).

    [91] H. G. Hedderich, R. Engleman, D. R. Huffman, C. I. Frum, P. F. Bernath, L. D. Lamb. The infrared emission spectrum of gas-phase C60 (buckmisterfullerene). Chem. Phys. Lett., 176, 504-508(1991).

    [92] E. Burkel, F. Zhang, F. Ahmed, C. Mihoc, C. Lathe. Thermal stability of carbon nanotubes, fullerene and graphite under spark plasma sintering. Chem. Phys. Lett., 510, 109-114(2011).

    [93] Y. Murata, K. Komatsu, G. Wang, M. Shiro. Synthesis and X-ray structure of dumb-bell-shaped C120. Nature, 387, 583-586(1997).

    [94] A. Soldatov, P. Nagel, C. Meingast, B. Sundqvist, V. Pasler, S. Lebedkin et al. C60 one- and two-dimensional polymers, dimers, and hard fullerite: Thermal expansion, anharmonicity, and kinetics of depolymerization. Phys. Rev. B, 60, 16920-16927(1999).

    [95] Y. Saito, Y. Ando, M. Kato, M. Ohkohchi, H. Shinohara, H. Nagashima. Electric conductivity and band gap of solid C60 under high pressure. Chem. Phys. Lett., 189, 236-240(1992).

    [96] M. Ata, M. Shiraishi, M. Ramm. The characterization of plasma-polymerized C60 thin films. Appl. Phys. A: Mater. Sci. Process., 74, 613-616(2002).

    [97] M. Tokumoto, P. Scharff, B. Narymbetov, H. Kobayashi, B. Sundqvist, T. L. Makarova et al. Anisotropic metallic properties of highly - oriented rhombohedral C60 polymer. Synth. Met., 121, 1099-1100(2001).

    [98] V. A. Davydov, T. L. Makarova, M. E. Gaevski, E. Olsson, B. Sundqvist, P. Scharff et al. Electrical properties of two-dimensional fullerene matrices. Carbon, 39, 2203-2209(2001).

    [99] L. G. Bulusheva, D. Tománek, A. V. Okotrub, T. L. Makarova, V. V. Belavin, V. A. Davydov. Electronic structure and properties of rhombohedrally polymerized C60. J. Chem. Phys., 115, 5637-5641(2001).

    [100] S. G. Buga, V. D. Blank, L. Edman, G. A. Dubitsky, E. B. Nyeanchi, X.-M. Zhu et al. Semimetallic and semiconductor properties of some superhard and ultrahard fullerites in the range 300–2 K. J. Phys. Chem. Solids, 61, 1009-1015(2000).

    [101] N. M. Fortunato, M. Melle-Franco, L. Marques, J. Laranjeira, M. Baroso, K. Strutyński. Three-dimensional C60 polymers with ordered binary-alloy-type structures. Carbon, 137, 511-518(2018).

    [102] L. Marques, M. Mezouar, J. Laranjeira, K. Strutyński, M. Melle-Franco. Bonding frustration in the 9.5 GPa fcc polymeric C60. Phys. Status Solidi Rapid Res. Lett., 11, 1700343(2017).

    [103] O. Andersson, A. Soldatov. Thermal conductivity of pressure polymerized C60. Appl. Phys. A, 64, 227-229(1997).

    [104] M. Yao, W. Cui, D. Liu, B. Liu, Q. Li, L. Wang et al. Pressure-induced phase transitions of C70 nanotubes. J. Phys. Chem. C, 115, 8918-8922(2011).

    [105] B. Sundqvist. Intermolecular bonding in C70 at high pressure and temperature. Carbon, 125, 258-268(2017).

    [106] H. Sato, H. Shinohara, Y. Akahama, H. Kawamura, M. Kobayashi, Y. Saito. Orientational ordering in solid C70 under high pressure. Solid State Commun., 83, 563-565(1992).

    [107] K. Prassides, T. J. S. Dennis, C. Christides, M. Thomas. Pressure and temperature evolution of the structure of solid C70. Europhys. Lett., 22, 611-618(1993).

    [108] S. Lebedkin, D. Johnels, M. Haluska, A. V. Soldatov, A. Dzyabchenko, G. Roth, H. Kuzmany, C. Meingast et al. Topochemical polymerization of C70 controlled by monomer crystal packing. Science, 293, 680-683(2001).

    [109] B. N. Mavrin, V. D. Blank, G. A. Dubitsky, N. R. Serebryanaya, V. N. Denisov, S. G. Buga et al. Polymerization and phase diagram of solid C70 after high-pressure-high-temperature treatment. Phys. Lett. A, 248, 415-422(1998).

    [110] O. M. Zhigalina, B. A. Kulnitskiy, V. D. Blank. Dimerisation and polymerisation of C70 after thermobaric. Carbon, 38, 2051-2054(2000).

    [111] R. Soares, L. Marques, Y. Skorokhod. Extended polymerization in ABC-stacked C70 fullerite. Carbon, 82, 599-603(2015).

    [112] R. Soares, Y. Skorokhod, L. Marques. A new fullerene network phase obtained from C70 at high-pressure and high-temperature. Phys. Status Solidi Rapid Res. Lett., 9, 535-538(2015).

    [113] A. Demortier, A. Fonseca, J. B. Nagy, R. Doome. Study of the anomalous solubility behavior of solutions saturated with fullerenes by 13C-NMR. Magn. Reson. Colloid Interface Sci., 76, 513-518(2002).

    [114] P. Marsh, T. Siegrist, R. M. Fleming, F. A. Thiel, A. R. Kortan, B. Hessen et al. Pseudotenfold symmetry in pentane-solvated C60 and C70. Phys. Rev. B, 44, 888-891(1991).

    [115] V. Agafonov, F. Michaud, M. Barrio, J. L1. Tamarit, D. O. López, S. Toscani et al. Solid-state studies on a C60 solvate grown from 1,1,2-trichloroethane. Chem. Mater., 12, 3595-3602(2000).

    [116] J. L1. Tamarit, D. O. López, R. Céolin, P. Espeau, M. Barrio. Solid-state studies of C60 solvates formed in the C60-BrCCl3 system. Chem. Mater., 15, 288-291(2003).

    [117] R. Céolin, D. O. López, P. Espeau, M. Barrio, H. Allouchi, J. LI. Tamarit et al. Solid state studies of the C60. 2(CH3)CCl3 solvate. Carbon, 43, 417-424(2005).

    [118] S. Toscani, M. Barrio, J. LI. Tamarit, D. O. López, V. Agafonov, H. Allouchi et al. Decagonal C60 crystals grown from n-hexane solutions: Solid-state and aging studies. Chem. Phys. Lett., 330, 491-496(2000).

    [119] R. D. Sherwood, S. M. Gorun, D. M. Cox, K. M. Creeqan, V. W. Day, C. S. Day et al. Solvated C60 and C60/C70 and the low-resolution single crystal X-ray structure of C60. J. Chem. Soc., Chem.Commun., 21, 1556-1558(1991).

    [120] Y. L. Slovokhotov, I. S. Neretin, A. V. Dzyabchenko, M. V. Korobov, A. L. Mirakyan, E. B. Stukalin et al. New solid solvates of C60 and C70 fullerenes: The relationship between structures and lattice energies. Carbon, 41, 2743-2755(2003).

    [121] H. Allouchi, J. LI. Tamarit, R. Céolin, D. O. López, M. Barrio, S. Toscani et al. Solid-state studies on a cubic 1:1 solvate of C60 grown from dichloromethane and leading to another hexagonal C60 polymorph. Chem. Mater., 13, 1349-1355(2001).

    [122] M. Barrio, H. Allouchi, D. O. López, R. Céolin, P. Espeau, J. L1. Tamarit et al. Phase equilibria in the C60+ ferrocene system and solid-state studies of the C60•2ferrocene solvate. Chem. Mater., 14, 321-326(2002).

    [123] P. Byszewski, R. Świetlik, E. Kowalska. Interactions of C60 with organic molecules in solvate crystals studied by infrared spectroscopy. Chem. Phys. Lett., 254, 73-78(1996).

    [124] A. Talyzin, U. Jansson. C60 and C70 solvates studied by Raman spectroscopy. J. Phys. Chem. B, 104, 5064-5071(2000).

    [125] M. Romanini, J. L1. Tamarit, R. Macovez, E. Mitsari, N. Qureshi, M. Barrio. C60 solvate with (1,1,2)-trichloroethane: Dynamic statistical disorder and mixed conformation. J. Phys. Chem. C, 120, 12831-12839(2016).

    [126] J. S. Schilling, A. K. Gangopadhyay, K. Robinson, T. Kowalewski, M. De. Leo, W. E. Buhro. Synthesis and characterization of C60{CCl4}10. Solid State Commun., 96, 597-600(1995).

    [127] R. Świetlik, A. Graja. Temperature study of IR spectra of some C60 compounds. Synth. Met., 70, 1417-1418(1995).

    [128] H. C. Choi, T. Joo, C. Park, E. Yoon, M. Kawano. Self-Crystallization of C70 cubes and remarkable enhancement of photoluminescence. Angew. Chem., Int. Ed., 49, 9670-9675(2010).

    [129] M. Yao, S. Chen, M. Du, X. Yang, R. Liu, B. Sundqvist et al. Effect of C70 rotation on the photoluminescence spectra of compressed C70∗mesitylene. J. Raman Spectrosc., 48, 437-442(2016).

    [130] S. Pekker, C. A. Kuntscher, G. Klupp, É. Kováts, K. Kamarás, S. Frank et al. Pressure-dependent infrared spectroscopy on the fullerene rotor-stator compound C60-C8H8. Phys. Status Solidi B, 243, 2981-2984(2006).

    [131] Q. Li, M. Yao, Z. Li, L. Wang, W. Cui, D. Liu et al. Synthesis and solid-state studies of self-assembled C60 microtubes. Diamond Relat. Mater., 20, 178-182(2011).

    [132] D. V. Konarev, K. P. Meletov. Raman study of the pressure-induced phase transitions in the molecular donor-acceptor complex {Pt(dbdtc)2}C60. Chem. Phys. Lett., 553, 21-25(2012).

    [133] K. P. Meletov, D. V. Konarev. Raman study of the pressure-induced charge transfer transition in the neutral donor-acceptor complexes {Ni(nPr2dtc)2}(C60)2 and {Cu(nPr2dtc)2}(C60)2. Fullerenes, Nanotubes, Carbon Nanostruct., 20, 336-340(2012).

    [134] H. Kawamura, Y. Akahama, M. Kobayashi, H. Sato, Y. Saito, H. Shinohara. Electrical resistance of iodine-doped C60 under high pressure. Solid State Commun., 82, 605-607(1992).

    [135] Q. Li, F. Ma, M. Yao, Z. Yao, R. Liu, W. Cui et al. Reversible pressure-induced polymerization of Fe(C5H5)2 doped C70. Carbon, 62, 447-454(2013).

    [136] I. Jalsovszky, C. A. Kuntscher, G. Klupp, K. Kamarás, B. J. Nagy, K. Thirunavukkuarasu et al. Orientational ordering and intermolecular interactions in the rotor-stator compounds C60•C8H8 and C70•C8H8 studied under pressure. J. Phys. Chem. C, 112, 17525-17532(2008).

    [137] U. Jansson, A. V. Talyzin, L. S. Dubrovinsky. High pressure Raman study of C60S16. Solid State Commun., 123, 93-96(2002).

    [138] T. Kawamoto, A. Omerzu, M. Tokumoto, H. Sakamoto, M. Takei, K. Mizoguchi et al. Uniaxial strain study in purely organic ferromagnet α-TDAE-C60-Mechanism and structure. Polyhedron, 24, 2173-2175(2005).

    [139] M. Machino, K. Mizoguchi, T. Kawamoto, M. Tokumoto, H. Sakamoto, A. Omerzu et al. Pressure effect in TDAE-C60 ferromagnet: Mechanism and polymerization. Phys. Rev. B, 63, 140417(2001).

    [140] D. Liu, W. Cui, R. Liu, B. Zou, Q. Li, M. Yao et al. Reversible polymerization in doped fullerides under pressure: The case of C60(Fe(C5H5)2)2. J. Phys. Chem. B, 116, 2643-2650(2012).

    [141] M. Tachibana, K. Kato, H. Gonnokami, H. Murata. Polymerization in ferrocene-doped C60 nanosheets under high pressure and light irradiation. Carbon, 107, 622-628(2016).

    [142] M. Du, J. Dong, M. Yao, É. Kováts, P. Ge, Q. Dong et al. New ordered structure of amorphous carbon clusters induced by fullerene-cubane reactions. Adv. Mater., 30, 1706916(2018).

    [143] M. Yao, S. Liu, X. Yang, J. Xiao, M. Du, W. Cui et al. Tailoring building blocks and their boundary interaction for the creation of new, potentially superhard, carbon materials. Adv. Mater., 27, 3962-3968(2015).

    [144] D. Liu, M. Yao, W. Cui, Z. Li, L. Wang, Q. Li et al. In situ Raman and photoluminescence study on pressure-induced phase transition in C60 nanotubes. J. Raman Spectrosc., 43, 737-740(2012).

    [145] B. W. Smith, M. Monthioux, D. E. Luzzi. Encapsulated C60 in carbon nanotubes. Nature, 396, 323-324(1998).

    [146] B. W. Smith, D. E. Luzzi. Formation mechanism of fullerene peapods and coaxial tubes: A path to large scale synthesis. Chem. Phys. Lett., 321, 169-174(2000).

    [147] S. Saito, S. Okada, A. Oshiyama. Energetics and electronic structures of encapsulated C60 in a carbon nanotube. Phys. Rev. Lett., 86, 3835-3838(2001).

    [148] L. A. Girifalco, M. Hodak. Systems of C60 molecules inside (10,10) and (15,15) nanotube: A Monte Carlo study. Phys. Rev. B, 68, 085405(2003).

    [149] M. Chorro, S. Rols, L. Noé, J. Cambedouzou, M. Monthioux, A. Iwasiewicz-Wabnig et al. Discriminated structural behaviour of C60 and C70 peapods under extreme conditions. Europhys. Lett., 79, 56003(2007).

    [150] M. Yao, X. Yang, X. Wu, K. Yang, S. Liu, S. Chen et al. Novel superhard sp3 carbon allotrope from cold-compressed C70 peapods. Phys. Rev. Lett., 118, 245701(2017).

    [151] Q. Li, A. R. Oganov, H. B. Wang, H. Wang, Y. Ma, Y. Wang et al. Superhard monoclinic polymorph of carbon. Phys. Rev. Lett., 102, 175506(2009).

    [152] A. Lappas, K. Tanigaki, K. Prassides, M. Kosaka, C. M. Brown, S. Margadonna et al. Superconductivity in LixCsC60 fullerides. Phys. Rev. B, 59, R6628(1999).

    [153] S. Margadonna, K. Prassides. Recent advances in fullerene superconductivity. J. Solid State Chem., 168, 639-652(2002).

    [154] T. W. Ebbesen, S. Saito, S. Tsai, Y. Kubo, K. Tanigaki, J. Mizuki et al. Superconductivity at 33 K in CsxRbyC60. Nature, 352, 222-223(1991).

    [155] G. Bortel, S. Pekker, P. W. Stephens, A. Jánossy, G. Faigel, M. Tegze et al. Polymeric fullerene chains in RbC60 and KC60. Nature, 370, 636-639(1994).

    [156] M. V. Fernandez-Serra, R. Poloni, S. Le. Floch, G. Montagnac, S. Pascarelli, D. Machon et al. High-pressure stability of Cs6C60. Phys. Rev. B, 77, 125413(2008).

    [157] S. Le. Floch, R. Poloni, P. Toulemonde, S. Pascarelli, G. Aquilanti, D. Machon et al. High-pressure phase transition in Rb6C60. Phys. Rev. B, 77, 205433(2008).

    [158] M. Riccò, M. Belli, D. Quintavalle, M. Mazzani, D. Pontiroli et al. Superionic Conductivity in the Li4C60 fulleride polymer. Phys. Rev. Lett., 102, 145901(2009).

    [159] M. Yao, T. Wågberg, B. Sundqvist. Effect of high pressure on electrical transport in the Li4C60 fulleride polymer from 100 to 400 K. Phys. Rev. B, 81, 155441(2010).

    [160] T. Wågberg, M. Yao, B. Sundqvist. Electrical transport properties of Na2C60 under high pressure. Phys. Rev. B, 80, 115405(2009).

    [161] V. N. Denisov, G. A. Dubitsky, V. D. Blank, A. N. Ivlev, N. R. Serebryanaya, S. G. Buga et al. Ultrahard and superhard carbon phases produced from C60 by heating at high pressure: Structural and Raman studies. Phys. Lett. A, 205, 208-216(1995).

    Cuiying Pei, Lin Wang. Recent progress on high-pressure and high-temperature studies of fullerenes and related materials[J]. Matter and Radiation at Extremes, 2019, 4(2): 28201
    Download Citation