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 31 毫秒
1.
Preparation of μ-Sulfurdisulfonium Salts [(CH3)2S? Sx? S(CH3)2]2+2A? (x = 1–3, A? = AsF6?, SbF6?, SbCl6?). On the Analogy of the Reactivity of Sulfanes and Sulfonium Salts The preparation of the μ-sulfurdisulfonium salts [(CH3)2S? Sx? S(CH3)2]2+(A?)2 with x = 1–3 and A? = AsF6?, SbF6?, SbCl6? is reported. The salts are formed by reaction of (CH3)2SH+A? and (CH3)2SSH+A? with SCl2 and S2Cl2, resp. They are characterized by vibrational spectroscopic measurements. [(CH3)2S? S2? S(CH3)2]2+(SbF6?)2 crystallizes in the space group C2/c with a = 1 884.5(7) pm, b = 1 302.8(5) pm, c = 1 477.2(5) pm, β = 98.62(3)° und Z = 8.  相似文献   

2.
The reactions of elemental nickel and tellurium and of ZnTe with excess AsF5 in liquid SO2 yield [M(SO2)6](Te6)[AsF6]6 (M = Ni, Zn) as orange crystals. The crystal structure determinations (triclinic, , M = Ni: a = 1632.59(2), b = 1795.06(1), c = 1822.97(2) pm, α = 119.11(4), β = 90.78(4), γ = 106.28(4)°, V = 4408.24(8)·106pm3, Z = 4) show the two compounds to be isotypic. The structures are made up of discrete [M(SO2)6]2+ complexes, Te64+ clusters and octahedral [AsF6]? ions. In the [M(SO2)6]2+ complexes the metal ions are surrounded octahedrally by six SO2 molecules bound via the O atoms. The Te64+ polycations are of trigonal prismatic shape with short Te–Te bonds within the triangular faces (270 pm) and long Te–Te bonds along the edges parallel to the pseudo C3 axes of the prisms (312 pm). The arrangement of the ions is related to the Li3Bi structure type. [M(SO2)6]2+ complexes and Te64+ polycations together form a distorted cubic closest packing with all tetrahedral and octahedral interstices filled by [AsF6]? ions. The analogous reaction starting from CdTe did not yield a compound containing simultaneously [Cd(SO2)n]2+ complexes and tellurium polycations but instead Te6[AsF6]4 · 2 SO2 besides [Cd(SO2)2][AsF6]2 were obtained. It crystallizes isotypically to [Mn(SO2)2][AsF6]2 (Mews, Zemva, 2001) (orthorhombic, Fdd2, a = 1534.96(3), b = 1812.89(3), c = 892.28(3) pm, V = 2483·106 pm3, Z = 4).  相似文献   

3.
Thermal decomposition of bis(hexamethylbenzene)iron(0) in the presence of carbon monoxide yields a novel carbonyl iron complex, [C6(CH3)6]Fe(CO)2. The cyclohexadiene complex [C6(CH3)6]Fe(C6H8) is obtained from reaction of bis(hexamethylbenzene)iron(0) with either 1,3-cyclohexadiene or benzene, and the yield is much greater in the presence of hydrogen gas. Interaction of bis-(hexamethylbenzene)iron(0) with 2-butyne induces a catalytic cyclotrimerization to give more hexamethylbenzene. Kinetic and isotope distribution studies indicate that the primary step in these reactions is not a direct loss of one ring ligand, but rather an insertion of the iron center into one of the ligand methyl CH bonds, leading to a benzyl hydride complex species. Mechanisms for the subsequent reactions of this iron hydride species are proposed.  相似文献   

4.
Pentafluorosulfanylamines and Sulfanylammonium Salts . From the addition of HF to sulfurtetrafluorideimides N-alkylpentafluorosulfanylamines RNHSF5(2a, 2b: R=CH3, C2H5) are obtained in quantitative yield. N, N-dialkylpentafluorosuIfanylamines Et2NSF5(5a) and pip-SF5 (5b) are isolated from the reaction of the appropriate sulfurdifluoronitridearnides NSF2NR2 and HF/SF4. Protonation of the amines with the superacidic system HF/AsF5 gives stable pentafluorosulfanyl-ammonium salts SF5NHRR′. AsF6 (12: R = R′ = CH3; 14: R=R′=H; 10: R=CH3, R′ = H). Under the same conditions the adduct AsF5· NSF2CF(CF3)2 (15) forms a cation with hexacoordinated sulfur (trans-H3NSF4CF(CF3)2?AsF66: 16), while with Asp5 · NSF2NMe2 (17) the reaction stops at tetracoordination (HNSF2NMe2+AsF6 : 18).  相似文献   

5.
A structurally diverse range of lipophilic, cationic η6‐arene η5‐cyclopentadienyl (η5‐Cp*) full‐sandwich complexes of ruthenium(II) have been prepared and structurally characterized by Fourier‐transform IR and NMR spectroscopy, electrospray mass spectrometry, and elemental microanalyses. Computational experiments incorporating the Hartree–Fock theory and the second‐order Møller–Plesset perturbation theory predict each complex to possess a uniform δ+ electrostatic potential, with the cationic charge of the [RuCp*]+ moiety completely delocalizing throughout the molecular structure of each metallocene. In vitro cytotoxicity studies demonstrate these delocalized lipophilic cations to be potent growth inhibitors of eleven unique tumorigenic cell lines, while exhibiting significantly lower levels of toxicity towards both a normal human fibroblast and a mouse macrophage cell line. Single‐crystal X‐ray structural determinations are additionally reported for five complexes, [Ru(η6‐C6H5(CH2)2CH3)(η5‐C5(CH3)5)]BPh4, [Ru(η6‐C6H5CO2CH2CH3)(η5‐C5(CH3)5)]BF4, [Ru(η6‐C10H8)(η5‐C5(CH3)5)]BPh4, [Ru(η6‐C14H10)(η5‐C5(CH3)5)]BPh4, and [Ru(η6‐C16H10)(η5‐C5(CH3)5)]BPh4.  相似文献   

6.
In contrast to the well‐known 2‐norbornyl cation, the structure of which was a matter of long debate until its pentacoordinated nature was recently proven by an X‐ray structure, the pentagonal‐pyramidal dication of hexamethylbenzene has received considerably less attention. This species was first prepared by Hogeveen in 1973 at low temperatures in magic acid (HSO3F/SbF5), for which he proposed a non‐classical structure (containing a hexacoordinated carbon) based on NMR spectroscopy and reactivity studies, but no X‐ray crystal structure has been reported. C6(CH3)62+ can be obtained through the dissolution of hexamethyl Dewar benzene epoxide in HSO3F/SbF5 and crystallized as the SbF6 salt upon addition of excess anhydrous hydrogen fluoride. The crystal structure of C6(CH3)62+ (SbF6)2⋅HSO3F confirms the pentagonal pyramidal structure of the dication. The apical carbon is bound to one methyl group (distance 1.479(3) Å) and to the five basal carbon atoms (distances 1.694(2)–1.715(3) Å).  相似文献   

7.
The preparation of the first niobium(V) and molybdenum(VI) dimethylmetallocene cations is reported. [Cp2Nb(CH3)2]+[AsF6] (1) and [Cp2Mo(CH3)2]2+ [AsF6]2 (2) (Cp = η5-C5H5) are prepared by oxidation of Cp2Nb(CH3)2 and Cp2Mo(CH3)2 with AsF5 in liquid sulfur dioxide. IR investigations confirm the ionic structure of 1 and 2, and that the AsF6 is not coordinated to the metal centre.  相似文献   

8.
Metal Complexes of Biologically Important Ligands, CLVII [1] Halfsandwich Complexes of Isocyanoacetylamino acid esters and of Isocyanoacetyldi‐ and tripeptide esters (?Isocyanopeptides”?) N‐Isocyanoacetyl‐amino acid esters CNCH2C(O) NHCH(R)CO2CH3 (R = CH3, CH(CH3)2, CH2CH(CH3)2, CH2C6H5) and N‐isocyanoacetyl‐di‐ and tripeptide esters CNCH2C(O)NHCH(R1)C(O)NHCH(R2)CO2C2H5 and CNCH2C(O)NHCH(R1)C(O)NHCH (R2)C(O)NHCH(R3)CO2CH3 (R1 = R2 = R3 = CH2C6H5, R2 = H, CH2C6H5) are available by condensation of potassium isocyanoacetate with amino acid esters or peptide esters. These isocyanides form with chloro‐bridged complexes [(arene)M(Cl)(μ‐Cl)]2 (arene = Cp*, p‐cymene, M = Ir, Rh, Ru) in the presence of Ag[BF4] or Ag[CF3SO3] the cationic halfsandwich complexes [(arene)M(isocyanide)3]+X? (X = BF4, CF3SO3).  相似文献   

9.
π-Arene complexes of cadmium(II) and zinc(II) have been prepared from the first time. The 1:1 complexes Cd(AsF6)2. Arene(Arene=hexaethylor hexamethylbenzene, pentamethylbenzene, durene, p-xylene or benzene), Cd(SbF6)2. Arene(Arene = hexamethylbenzene, toluene or benzene) and Zn(SbF6)2. Arene(Arene = hexamethylbenzene or pentamethylbenzene) are synthesized from the strong acid salt and arene in liquid sulfur dioxide. 1H and 13C NMR spectra are consistent with localized bonding of the arene to the metal cation. Exchange-averaged vn]13C chemical shifts for the systems Cd(AsF6)2-arene-SO2 confirm the 1:1 stoichiometry in solution and suggest that the stabilities of the complexes are in the approximate range 0.48 – 2.1 M?1 for the series benzene-hexamethylbenzene. For the system Cd(AsF6)2-C6Me6-SO2, a detailed 113Cd NMR study is consistent with the solution stoichiometry and stability determined from 13C NMR. In general, complexation to an arene produces deshielding of the 113Cd resonance of Cd(AsF6)2.  相似文献   

10.
The absolute bimolecular rate constants for the reactions of C6H5 with 2‐methylpropane, 2,3‐dimethylbutane and 2,3,4‐trimethylpentane have been measured by cavity ringdown spectrometry at temperatures between 290 and 500 K. For 2‐methylpropane, additional measurements were performed with the pulsed laser photolysis/mass spectrometry, extending the temperature range to 972 K. The reactions were found to be dominated by the abstraction of a tertiary C H bond from the molecular reactant, resulting in the production of a tertiary alkyl radical: C6H5 + CH(CH3)3 → C6H6 + t‐C4H9 (1) (1) C6H5 + (CH3)2CHCH(CH3)2 → C6H6 + t‐C6H13 (2) (2) C6H5 + (CH3)2CHCH(CH3)CH(CH3)2 → C6H6 + t‐C8H17 (3) (3) with the following rate constants given in units of cm3 mol−1 s−1: k1 = 10(11.45 ± 0.18) e−(1512 ± 44)/T k2 = 10(11.72 ± 0.15) e−(1007 ± 124)/T k3 = 10(11.83 ± 0.13) e−(428 ± 108)/T © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 645–653, 1999  相似文献   

11.
The Donor Properties of Bis(pyrazolyl)‐Sulfur Derivatives From the reactions of bis(pyrazolyl)sulfane S(pz)2 ( 1 ) with the fluoro Lewis acids BF3 and AsF5 in liquid SO2 the 1:2‐adducts S(pz·BF3)2 ( 2 ) and S(pz·AsF5)2 ( 3 ) are obtained. 1 reacts with [Co(SO2)4(FAsF5)2] to give the doubly bridged FAsF4F dimeric complex [Co{S(pz)2}(FAsF5)(SO2)(μ‐FAsF4F)]2 ( 5 ). From F2S(pz)2 and [Ni(SO2)6](AsF6)2, the fluorocubane [Ni4F4{S(pz)2}4(μ‐FAsF4F)2](AsF6)2·4SO2 ( 8 ) is isolated. The X‐ray structures of the compounds 2 , 3 , 5 and 8 are reported.  相似文献   

12.
Abstract : γ-Butyrolactone and γ-butyrolactam were reacted in the superacidic systems XF/MF5 (X=H, D; M=As, Sb). Salts of the monoprotonated species of γ-butyrolactone were obtained in terms of [(CH2)3OCOH]+[AsF6], [(CH2)3OCOH]+[SbF6] and [(CH2)3OCOD]+[AsF6] and the analogous lactam salts in terms of [(CH2)3NHCOH]+[AsF6], [(CH2)3NHCOH]+[SbF6] and [(CH2)3NDCOD]+[AsF6]. The salts were characterized by low temperature Raman and infrared spectroscopy and for both protonated hexafluoridoarsenates, [(CH2)3OCOH]+[AsF6] and [(CH2)3NHCOH]+[AsF6], single-crystal X-ray structure analyses were conducted. In addition to the experimental results, quantum chemical calculations were performed on the B3LYP/aug-cc-pVTZ level of theory. As in both crystal structures C⋅⋅⋅F contacts were observed, the nature of these contacts is discussed with Mapped Electrostatic Potential as a rate of strength.  相似文献   

13.
Hg2(CH3SO3)2: Synthesis, Crystal Structure, Thermal Behavior, and Vibrational Spectroscopy Colorless single crystals of Hg2(CH3SO3)2 are formed in the reaction of HgO, Hg, and HSO3CH3. In the monoclinic compound (I2/a, Z = 4, a=883.2(2), b=854.0(2), c=1188.9(2) pm, β = 92.55(2)°, Rall=0.0445) the Hg22+ ion is coordinated by two monodentate CH3SO3 anions. Further contacts Hg‐O occur in the range from 262 to 276 pm and lead to a linkage of the [Hg2(CH3SO3)2] units. The thermal analysis shows that Hg2(CH3SO3)2 decomposes at 300° yielding elemental mercury. The mass numbers of the species evolved lead to the assumtion that SO3, SO2, CO2, CO and H2CO are formed during the reaction. In the IR and the Raman spectrum the typical vibrations of the CH3SO3 ion are observed, the Raman spectrum shows the Hg‐Hg stretching vibration at 177 cm—1 within the Hg22+ ion additionally.  相似文献   

14.
Report of the preparation, chemical properties, and the infrared-to-ultra-violet spectra of the perchlorates and bromides of the two complex cations [Co2{ac(OH)2}(NH3)6]3+ (where ac = HCO2, CH3CO2; CH2ClCO2, CHCl2CO2, CCl3CO2, CHFCO2, CHF2CO2 und CF3CO2) and [Co2{ac2(OH)}(NH3)6]3+ (where ac = CH2ClCO2, CHClCO2 und CCl3CO2). The perchlorate, nitrate, bromide and dithionate salts of the tetranuclear complex [Co4{C2O4(OH)4}(NH3)12]6+ are described. The complex reported by WERNER as [Co2{OH}2(CH3CO2)H2O(NH3)6]Br3 actually has the formula [Co2{CH3CO2(OH)2}(NH3)6]Br3 · CH3COOH.  相似文献   

15.
The mechanism of the reactions of electronically excited SO2 with isobutane has been studied through the measurement of the initial quantum yields of product formation in 3130 Å irradiated gaseous binary mixtures of SO2 and isobutane and ternary mixtures of SO2, isobutane, C6H6 or CO2. Under low-pressure conditions (P < 10 torr) the kinetic treatment of the present data shows that only one singlet and one triplet state, presumably the 1B1 and 3B1 states, are involved in the photoreaction mechanism. The data give k2a = 8.4 × 109; SO2(1B1) + isobutane → products (2a); k5a ? k5 = 8.7 × 108 l./mol·sec; SO2(3B1) + isobutane → products (5a) SO2(3B1) + isobutane → (SO2) + isobutane (5b) k1a/k1 = 0.145 ± 0.037; SO2(1B1) + SO2 → SO2(3B1) + SO2 (1a) SO2(1B1) + SO2 → (2SO2) (1b) k2b/k2 = 0.273 ± 0.018; SO2(1B1) + isobutane → SO2(3B1) + isobutane (2b); SO2(1B1) + isobutane → (SO2) + isobutane (2c) error limits are ± 2 σ. The contribution from the excited SO2(1B1) molecules to the quantum yields of the photolyses of SO2–isobutane mixtures is not negligible. Under high-pressure conditions (P > 10 torr) the low-pressure mechanism coupled with the saturation effect on the phosphorescence lifetimes of SO2(3B1) molecules cannot alone rationalize the quantum yields. The evaluation suggests that some nonradiative intermediate state (X) is involved in the formation of “extra” triplet molecules. This ill-defined state decays largely nonradiatively to SO2 in experiments at low pressures, X → SO2 (12). In the presence of C6H6 the low-pressure data give k7 = (8.5 ± 1.8) × 1010, and the high-pressure data give k7 = (8.3 ± 0.6) × 1010 and (9.9 ± 0.9) × 1010l./mol·sec; SO2(3B1) + C6H6 → nonradiative products (7). These estimates are in good agreement with values directly measured from low-pressure lifetime studies, (8.1 ± 0.7) × 1010 and (8.8 ± 0.8) × 1010l./mol·sec.  相似文献   

16.
The reaction of graphite fluorosulfate with an excess of arsenic(V) fluoride produces the new intercalation compound C14+ [AsF5(SO3F)]?. Identification as a first stage compound rests on the observed interlayer separation of 7.92 Å and the position of νE2g at 1636 cm?1 in the Raman spectrum. Evidence for [AsF5(SO3F)]? as intercalant is based on 19F NMR spectra of the solid material. C14+[AsF5(SO3F)]? is found to exhibit high electrical conductivities in the basal plane.  相似文献   

17.
The reaction of Cu(ClO4)2·6H2O, NaAsF6 and excess pyrazole yields hexakis­(pyrazole‐κN2)copper(II) bis­(hexa­fluoroarsenate), [Cu(C3H4N2)6](AsF6)2 or [Cu(pzH)6](AsF6)2 (pzH is pyrazole), (I). The analogous hexakis­(pyrazole‐κN2)copper(II) hexafluorophosphate perchlorate complex, [Cu(C3H4N2)6](PF6)1.29(ClO4)0.71 or [Cu(pzH)6](PF6)1.29(ClO4)0.71, (II), is obtained in a similar fashion, using KPF6 in place of NaAsF6. Both compounds contain the hitherto unknown [Cu(pzH)6]2+ complex cation, in which the copper(II) ion lies at the center of a regular octahedron of coordinated N atoms. The cation has crystallographically imposed symmetry. The X‐ray data indicate that the lack of the expected distortion can be accounted for by the presence of either static Jahn–Teller disorder or dynamic Jahn–Teller distortion.  相似文献   

18.
Graphite intercalated by AsF5 has been reported to give compounds of formula C8nAsF5 where n is the stage. It is doubtful however if materials of exact composition C8nAsF5 have ever been obtained. The intercalation of graphite by AsF5 is associated with electron oxidation of the graphite according to the equation: 3AsF5 + 2e? → 2AsF6? + AsF3. Because of the easy removal or displacement of AsF3 the As:F ratio is readily increased beyond 5. By treating graphite with excess AsF5, removing volatiles under vacuum and repeating the cycle seven times a first stage salt C10+AsF6? (Co = 7.96 ā) is made. Interaction of graphite with AsFs in the molar ratio 8:1, within a small volume reactor, yields a material of approximate composition C8AsF5. The major component of the volatiles at the onset of their removal is AsF5,, but, at a composition close to C10AsF5, is AsF3. ‘Graphite AsF5’ can be prepared by adding AsF3 to CxAsF6 salts. The electrical conductivities of ‘AsF5’ and AsF6 relatives will be compared and discussed.  相似文献   

19.
Synthesis, Crystal Structure, and Solid State Phase Transition of Te4[AsF6]2·SO2 The oxidation of tellurium with AsF5 in liquid SO2 yields Te42+[AsF6]2 which can be crystallized from the solution in form of dark red crystals as the SO2 solvate. The crystals are very sensitive against air and easily lose SO2, so handling under SO2 atmosphere or cooling is required. The crystal structure was determined at ambient temperature, at 153 K, and at 98 K. Above 127 K Te4[AsF6]2·SO2 crystallizes orthorhombic (Pnma, a = 899.2(1), b = 978.79(6), c = 1871.61(1) pm, V = 1647.13(2)·106pm3 at 297 K, Z = 4). The structure consists of square‐planar Te42+ ions (Te‐Te 266 pm), octahedral [AsF6] ions and of SO2 molecules which coordinate the Te4 rings with their O atoms in bridging positions over the edges of the square. At room temperature one of the two crystallographically independent [AsF6] ions shows rotational disorder which on cooling to 153 K is not completely resolved. At 127 K Te4[AsF6]2·SO2 undergoes a solid state phase transition into a monoclinic structure (P1121/a, a = 866.17(8), b = 983.93(5), c = 1869.10(6) pm, γ = 96.36(2)°, V = 1554, 2(2)·106 pm3 at 98 K, Z = 4). All [AsF6] ions are ordered in the low temperature form. Despite a direct supergroup‐subgroup relationship exists between the space groups, the phase transition is of first order with discontinuous changes in the lattice parameters. The phase transition is accompanied by crystal twinning. The main difference between the two structures lies in the different coordination of the Te42+ ion by O and F atoms of neighbored SO2 and [AsF6] molecules.  相似文献   

20.
CF3SO2N?SCl2 reagiert mit (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 oder (C5H10)S[NSi(CH3)3]2 unter Trimethylchlorsilanabspaltung zu den achtgliedrigen S4N4-Derivaten S4N4(NSO2CF3)2(CH3)4 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In den achtgliedrigen SN-Ringen haben die Schwefelatome die Koordinationszahl 3 und 4. Die Röntgenstrukturanalyse von 4a ergab eine Sessel-Konformation. 4a kristallisiert orthorhombisch in der Raumgruppe Pna21 mit a = 17,641(4), b = 6,406(2), c = 19,130(4) Å, dx = 1,815 g cm?3 und Z = 4. Die mittleren S? N-Abstände betragen an den vierfach koordinierten Schwefelatomen 1,597 Å und an den Schwefelatomen mit der Koordinationszahl 3 1,650 Å. CF3SO2N? SCl2 reagiert mit trimethylzinnhaltigen S? N-Verbindungen zum bekannten CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 und Dimethylzinndichlorid. Synthesis and X-Ray Structure Analysis of S4N4-Derivatives with Threefold and Fourfold Coordinated Sulfur Atoms CF3SO2N?SCl2 reacts with (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 or (C5H10S[NSi(CH3)2]2 under elimination of (CH3)3SiCl to yield the eight-membered S4N4 derivatives S4N4?NSO2CF3)2(CH3)4, 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In the eight-membered SN-rings the sulfur atoms have the coordination number 3 and 4. The X-ray structure analysis of 4a revealed a chair conformation. 4a crystallizes in the orthorhombic space group Pna21 with a = 17.641(4), b = 6.406(2), c = 19.130(4) Å, dx = 1.815 g cm?3, and Z = 4. The average S? N distance was found to be 1.597 Å at fourfold coordinated sulfur atoms and 1.650 Å at sulfur with coordination number 3. CF3SO2N=SCl2 reacts with trimethyl tin-containing S? N compounds to the known CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 and dimethyl tin dichloride.  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