首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
《Mendeleev Communications》2023,33(4):503-504
Chlorofullerenes C86(16)Cl16 and C86(17)Cl18,20 were prepared by chlorination of Cs-C86(16) and C2-C86(17), respectively, with VCl4 at 320–340 °C. An X-ray crystallo- graphic study with the use of synchrotron radiation revealed the chlorination patterns which show certain similarity due to only small differences between isomeric C86 cages.  相似文献   

2.
High‐temperature chlorination of fullerene C88 (isomer 33) with VCl4 gives rise to skeletal transformations affording several nonclassical (NC) fullerene chlorides, C86(NC1)Cl24/26 and C84(NC2)Cl26, with one and two heptagons, respectively, in the carbon cages. The branched skeletal transformation including C2 losses as well as a Stone–Wales rearrangement has been comprehensively characterized by the structure determination of two intermediates and three final chlorination products. Quantum‐chemical calculations demonstrate that the average energy of the C?Cl bond is significantly increased in chlorides of nonclassical fullerenes with a large number of chlorinated sites of pentagon–pentagon adjacency.  相似文献   

3.
High‐temperature chlorination of a fullerene C86 with VCl4 afforded non‐classical C84Cl30 and C82Cl30 containing one and two heptagons, respectively, in the carbon cages. Two types of C2 losses, which differ in the final arrangements of separate or fused pentagons, can occur successively in either order, producing rather flat or concave regions on the shrinked carbon cage. In the chlorination‐promoted skeletal transformation of C86 (isomer no. 16) with the loss(es) of C2 units, the structures of the starting, intermediate, and final compounds were all revealed unambiguously by X‐ray single crystal diffraction.  相似文献   

4.
The chlorination of HPLC fractions with pristine giant fullerenes, C102 and C104, followed by X‐ray crystallographic study of chlorides, C102(603)Cl18/20 and C104(234)Cl16–22, confirmed the presence of the most stable IPR (IPR=Isolated Pentagon Rule) isomers, C102(603) and C104(234), in the fullerene soot. The discussion concerns the chlorination patterns of polychlorides and relative stability of pristine isomers of C102 and C104 fullerenes.  相似文献   

5.
Chlorination of a mixture of C86 isomers no. 16 (Cs) and no. 17 (C2) with VCl4 or a (TiCl4+Br2) mixture afforded crystalline chlorides with 16 to 22 Cl atoms per fullerene cage. Single crystal X‐ray diffraction with the use of synchrotron radiation enabled us to determine the chlorination patterns of C86(16)Cl16, C86(17)Cl18, C86(17)Cl20, and C86(17)Cl22. At these degrees of chlorination, addition patterns of C86(16) and C86(17) chlorides have some features in common, owing to the close similarity in the cage structures of both isomers. The average energy of C?Cl bonds decreases with increasing number of attached Cl atoms.  相似文献   

6.
High‐temperature chlorination of three IPR isomers of fullerene C88, C2‐C88(7), Cs‐C88(17), and C2‐C88(33), resulted in the isolation and X‐ray structural characterization of C88(7)Cl12, C88(7)Cl24, C88(17)Cl22, and C88(33)Cl12/14. Chlorination patterns of C88(7) and C88(33) isomers are unusual in that one or more pentagons remain free from chlorination while some other pentagons are occupied by two or three Cl atoms. The addition patterns of the isolated chlorides are discussed in terms of the distribution of twelve pentagons on the carbon cages and the formation of stabilizing isolated C=C bonds and benzenoid rings.  相似文献   

7.
《化学:亚洲杂志》2017,12(18):2379-2382
Cage transformations in fullerenes are rare phenomena which are still not fully understood. We report the first skeletal transformation of an Isolated‐Pentagon‐Rule (IPR) isomer of C78 fullerene upon high‐temperature chlorination which proceeds by six‐step Stone–Wales rearrangements affording non‐IPR, non‐classical (NC ) C78(NC 2)Cl24 with two cage heptagons, six pairs of fused pentagons, and an unprecedented loop‐like chlorination pattern. The following loss of a C2 unit results in C76(NC 3)Cl24 containing three cage heptagons.  相似文献   

8.
High‐temperature chlorination of C90‐containing fullerene fraction resulted in the isolation and X‐ray structural characterization of C90(1)Cl10/12, the first derivatives of a relatively unstable isomer D5h‐C90(1) with a nanotubular shape. In the crystal structure, three isomers of both C90(1)Cl10 and C90(1)Cl12 with similar chlorination patterns co‐crystallize in the same crystallographic site. Thus, in contrast to the previous reports, D5h‐C90(1) is present, though with a low abundance, in the fullerene soot produced by arc‐discharge method with undoped graphite rods.  相似文献   

9.
High‐temperature chlorination of pristine C98 fullerene isomers separated by HPLC from the fullerene soot afforded crystals of C98Cl22 and C98Cl20. An X‐ray structure elucidation revealed, respectively, the presence of carbon cages of the most stable C2‐C98(248) and rather unstable C1‐C98(116), which represent the first isolated pentagon rule (IPR) isomers of fullerene C98 confirmed experimentally. The chlorination patterns of the chlorides are discussed in terms of the formation of isolated C=C bonds and aromatic substructures on the fullerene cages.  相似文献   

10.
Isolation and characterization of very large fullerenes is hampered by a drastic decrease of their content in fullerene soot with increasing fullerene size and a simultaneous increase of the number of possible IPR (Isolated Pentagon Rule) isomers. In the present work, fractions containing mixtures of C102 and C104 were isolated in very small quantities (several dozens of micrograms) by multi‐step recycling HPLC from an arc‐discharge fullerene soot. Two such fractions were used for chlorination with a VCl4/SbCl5 mixture in glass ampoules at 350–360 °C. The resulting chlorides were investigated by single‐crystal X‐ray diffraction using synchrotron radiation. By this means, two IPR isomers of C104, numbers 258 and 812 (of 823 topologically possible isomers), have been confirmed for the first time as chlorides, C1‐C104(258)Cl16 and D2‐C104(812)Cl24, respectively, while an admixture of C2‐C104(811)Cl24 was assumed to be present in the latter chloride. DFT calculations showed that pristine C104(812) belongs to rather stable C104 cages, whereas C104(258) is much less stable.  相似文献   

11.
Trifluoromethylation of higher fullerene mixtures with CF3I was performed in ampoules at 400 to 420 and 550 to 560 °C. HPLC separation followed by crystal growth and X‐ray diffraction studies allowed the structure elucidation of nine CF3 derivatives of D2‐C84 (isomer 22). Molecular structures of two isomers of C84(22)(CF3)12, two isomers of C84(22)(CF3)14, four isomers of C84(22)(CF3)16, and one isomer of C84(22)(CF3)20 were discussed in terms of their addition patterns and relative formation energies. DFT calculations were also used to predict the most stable molecular structures of lower CF3 derivatives, C84(22)(CF3)2–10. It was found that the addition of CF3 groups to C84(22) is governed by two rules: additions can only occur at para positions of C6(CF3)2 hexagons and no additions can occur at triple‐hexagon‐junction positions on the fullerene cage.  相似文献   

12.
Perfluoroalkylation of a higher fullerene mixture with CF3I or C2F5I, followed by HPLC separation of CF3 and C2F5 derivatives, resulted in the isolation of several C84(RF)n (n=12, 16) compounds. Single‐crystal X‐ray crystallography with the use of synchrotron radiation allowed structure elucidation of eight C84(RF)n compounds containing six different C84 cages (the number of the C84 isomer is given in parentheses): C84 (23)(C2F5)12 ( I ), C84 (22)(CF3)16 ( II ), C84 (22)(C2F5)12 ( III ), C84 (11)(C2F5)12 ( IV ), C84 (16)(C2F5)12 ( V ), C84 (4)(CF3)12 ( VI with toluene and VII with hexane as solvate molecules), and C84 (18)(C2F5)12 ( VIII ). Whereas some connectivity patterns of C84 isomers (22, 23, 11) had previously been unambiguously confirmed by different methods, derivatives of C84 isomers numbers 4, 16, and 18 have been investigated crystallographically for the first time, thus providing direct proof of the connectivity patterns of rare C84 isomers. General aspects of the addition of RF groups to C84 cages are discussed in terms of the preferred positions in the pentagons under the formation of chains, pairs, and isolated RF groups.  相似文献   

13.
High‐temperature chlorination of C100 fullerene followed by X‐ray structure determination of the chloro derivatives enabled the identification of three isomers of C100 from the fullerene soot, specifically numbers 18, 425, and 417, which obey the isolated pentagon rule (IPR). Among them, isomers C1‐C100(425) and C2‐C100(18) afforded C1‐C100(425)Cl22 and C2‐C100(18)Cl28/30 compounds, respectively, which retain their IPR cage connectivities. In contrast, isomer C2v‐C100(417) gives Cs‐C100(417)Cl28 which undergoes a skeletal transformation by the loss of a C2 fragment, resulting in the formation of a nonclassical (NC) C1‐C98(NC)Cl26 with a heptagon in the carbon cage. Most probably, two nonclassical C1‐C100(NC)Cl18/22 chloro derivatives originate from the IPR isomer C1‐C100(382), although both C1‐C100(344) and even nonclassical C1‐C100(NC) can be also considered as the starting isomers.  相似文献   

14.
Trifluoromethylation of a higher fullerene mixture with CF3I was performed in ampoules at 550 °C. HPLC separation followed by crystal growth and X‐ray diffraction study resulted in the structure elucidation of nine CF3 derivatives of D2d‐C84 (isomer 23). The molecular structures of C84(23)(CF3)4, C84(23)(CF3)8, C84(23)(CF3)10, C84(23)(CF3)12, two isomers of C84(23)(CF3)14, two isomers of C84(23)(CF3)16, and C84(23)(CF3)18 were discussed in terms of their addition patterns and the relative formation energies. Extensive theoretical DFT calculations were performed to identify the most stable molecular structures. It was found that the addition of CF3 groups to the C84(23) fullerene is governed by two main rules: no additions in positions of triple hexagon junctions and predominantly para additions in C6(CF3)2 hexagons on the fullerene cage. The only exception with an isolated CF3 group in C84(23)(CF3)12 is discussed in more detail.  相似文献   

15.
The carbon cage of buckminsterfullerene Ih-C60, which obeys the Isolated-Pentagon Rule (IPR), can be transformed to non-IPR cages in the course of high-temperature chlorination of C60 or C60Cl30 with SbCl5. The non-IPR chloro derivatives were isolated chromatographically (HPLC) and characterized crystallographically as 1809C60Cl16, 1810C60Cl24, and 1805C60Cl24, which contain, respectively two, four, and four pairs of fused pentagons in the carbon cage. High-temperature trifluoromethylation of the chlorination products with CF3I afforded a non-IPR CF3 derivative, 1807C60(CF3)12, which contains four pairs of fused pentagons in the carbon cage. Addition patterns of non-IPR chloro and CF3 derivatives were compared and discussed in terms of the formation of stabilizing local substructures on fullerene cages. A detailed scheme of the experimentally confirmed non-IPR C60 isomers obtained by Stone–Wales cage transformations is presented.  相似文献   

16.
Chlorination of the C100(18) fullerene with a mixture of VCl4 and SbCl5 gives rise to branched skeletal transformations affording non‐classical (NC) C94(NC1)Cl22 with one heptagon in the carbon cage together with the previously reported C96(NC3)Cl20 with three heptagons. The three‐step pathway to C94(NC1)Cl22 starts with two successive C2 losses of 5:6 C?C bonds to give two cage heptagons, whereas the third C2 loss of the 5:5 C?C bond from a pentalene fragment eliminates one of the heptagons. Quantum‐chemical calculations demonstrate that the two unusual skeletal transformations—creation of a heptagon in C96(NC3)Cl20 through a Stone–Wales rearrangement and the presently reported elimination of a heptagon through C2 loss—are both characterized by relatively low activation energy.  相似文献   

17.
Chlorination of various HPLC fractions of C96 with a mixture of VCl4 and SbCl5 at 340–360 °C and single‐crystal X‐ray diffraction study of the products led to the identification of three new IPR isomers of C96. The C96(175) isomer forms a stable chloride, C96(175)Cl20, while chlorides of two other new isomers, C96(114) and C96(80), undergo cage shrinkage yielding C94(NC1)Cl28 and C96(NC2)Cl32 with non‐classical (NC) cages. These two NC chlorides contain, respectively, one and two heptagons flanked by pairs of fused pentagons and are stabilized by chlorine attachment to the emerging pentagon–pentagon junctions. Thus, the number of the experimentally confirmed C96 isomers has reached nine, which corroborates the empirical rule that the C6n fullerenes exhibit particularly rich isomerism.  相似文献   

18.
X ray photoemission spectra show ~10 valence band features in C60 and C84 films, although they are less distinct in C84 because of the increased number of valence electrons and the decreased molecular symmetry. The presence of inequivalent carbon atoms in C84 is reflected by the fact that the C 1s main line is ~0.2 eV wider than in C60. The C 1s satellite region of C84 exhibits four features within 15 eV of the main line and shows that the HOMO-LUMO π-π* excitation energy of C84 is ~0.5 eV smaller than in C60. Potassium 2p spectra for K-doped C84 films suggest the formation of a phase with composition K3C84 with two spectroscopically-distinguishable K sites. A new spectral feature emerges when the K content is increased beyond K3C84, suggesting a phase transition and a saturated composition of K6C84, as for KxC60. These similarities are intriguing since KxC84 is insulating for all x while K3C60 is metallic and superconducting.  相似文献   

19.
The kinetics of the C2H5 + Cl2, n‐C3H7 + Cl2, and n‐C4H9 + Cl2 reactions has been studied at temperatures between 190 and 360 K using laser photolysis/photoionization mass spectrometry. Decays of radical concentrations have been monitored in time‐resolved measurements to obtain reaction rate coefficients under pseudo‐first‐order conditions. The bimolecular rate coefficients of all three reactions are independent of the helium bath gas pressure within the experimental range (0.5–5 Torr) and are found to depend on the temperature as follows (ranges are given in parenthesis): k(C2H5 + Cl2) = (1.45 ± 0.04) × 10?11 (T/300 K)?1.73 ± 0.09 cm3 molecule?1 s?1 (190–359 K), k(n‐C3H7 + Cl2) = (1.88 ± 0.06) × 10?11 (T/300 K)?1.57 ± 0.14 cm3 molecule?1 s?1 (204–363 K), and k(n‐C4H9 + Cl2) = (2.21 ± 0.07) × 10?11 (T/300 K)?2.38 ± 0.14 cm3 molecule?1 s?1 (202–359 K), with the uncertainties given as one‐standard deviations. Estimated overall uncertainties in the measured bimolecular reaction rate coefficients are ±20%. Current results are generally in good agreement with previous experiments. However, one former measurement for the bimolecular rate coefficient of C2H5 + Cl2 reaction, derived at 298 K using the very low pressure reactor method, is significantly lower than obtained in this work and in previous determinations. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 614–619, 2007  相似文献   

20.
Ternary Chlorides with Trigonal-Bipyramidal Clusters: [M5(C2)]Cl9 (M = La? Pr) The chlorides [M5(C2)]Cl9 (M = La? Pr) are obtained by metallothermic reduction of the respective trichlorides MCl3 with caesium in the presence of the lanthanide metal and carbon in sealed niobium ampoules at 800°C. They contain trigonal-bipyramidal clusters [M5(C2)] crystallizing with the triclinic crystal system. Only seven of the nine edges of the trigonal bipyramids are brigded by chloride (Cli). Each cluster is surrounded by twelve terminal ligands (Cla) so that units of the composition [M5(C2)Cl7i]Cl12a have to be considered. These are connected not only via Cli–a and Cla–a–a bridges. Rather, Cla–a (one linear and one bent) and Cli–i bridges are also observed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号