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.
Thiazyltrifluoride NSF3 and Thiazyldifluoridedimethylamide NSF2NMe2: Ligands in Organometallic Chemistry From the reaction of [Re(CO)5SO2]+AsF6? ( 1 ) and [CpFe(CO)2SO2]+AsF6? ( 6 ) with NSF3 ( 2 ) and NSF2NMe2 ( 4 ) the complexes [Re(CO)5NSF3]+AsF6? ( 3 ), [Re(CO)5NSF2NMe2]+AsF6? ( 5 ), [CpFe(CO)2NSF3]+AsF6? ( 7 ), and [CpFe(CO)2NSF2NMe2]+AsF6? ( 8 ) were obtained. The compounds have been characterised by X-ray crystallography, the ligand properties of 2 and 4 are discussed.  相似文献   

2.
During our studies towards the preparation of the pentagonal‐pyramidal hexamethylbenzene dication C6(CH3)62+, we isolated the unprecedented dicationic species C6(CH3)6SO2+ (AsF6)2 from the reaction of hexamethylbenzene with a mixture of anhydrous HF, AsF5, and liquid SO2. This compound can be understood as a complex of unknown SO2+ with hexamethylbenzene. Herein, we report on its synthesis, molecular structure, and spectroscopic characterization.  相似文献   

3.
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   

4.
A single crystal of Br3+AsF6? was isolated from a sample of BrF2+AsF6? which had been stored for 20 years. It was characterized by x-ray diffraction and Raman spectroscopy. It is shown that Br3+AsF6? (triclinic, a = 7.644(7) Å, b = 5.641(6) Å, c = 9.810(9) Å, α = 99.16(8)°, β = 86.61(6)°, γ = 100.11(7)°, space group P1 R(F) = 0.0608) is isomorphous with I3+AsF6?. The structure consists of discrete Br3+ and AsF6? ions with some cation-anion interaction causing distortion of the AsF6? octahedron. The Br3+ cation is symmetric with a bond distance of 2.270(5) Å and a bond angle of 102.5(2)°. The three fundamental vibrations of Br3+ were observed at 297 (ν3), 293 (ν1), and 124 cm?12). The Raman spectra of Cl3+AsF6? and I3+AsF6? were reinvestigated and ν3(B1) of I3+ was reassigned. General valence force fields are given for the series Cl3+, Br3+, and I3+. Reactions of excess Br2 with either BrF2+AsF6? or O2+AsF6? produce mixtures of Br3+AsF6? and Br5+AsF6?. Based on its Raman spectra, the Br5+ cation possesses a planar, centrosymmetric structure of C2h symmetry with three semi-ionically bound, collinear, central Br atoms and two more covalently, perpendicularly bound, terminal Br atoms.  相似文献   

5.
A reliable synthesis of unstable and highly reactive BrO2F is reported. This compound can be converted into BrO2+SbF6?, BrO2+AsF6?, and BrO2+AsF6??2 BrO2F. The latter decomposes into mixed‐valent Br3O4?Br2+AsF6? with five‐, three‐, one‐, and zero‐valent bromine. BrO2+ H(SO3CF3)2? is formed with HSO3CF3. Excess BrO2F yields mixed‐valent Br3O6+OSO3CF3? with five‐ and three‐valent bromine. Reactions of BrO2F and MoF5 in SO2ClF or CH2ClF result in Cl2BrO6+Mo3O3F13?. The reaction of BrO2F with (CF3CO)2O and NO2 produces O2Br‐O‐CO‐CF3 and the known NO2+Br(ONO2)2?. All of these compounds are thermodynamically unstable.  相似文献   

6.
Fluorination of Cyanuric Chloride and Low-Temperature Crystal Structure of [(ClCN)3F]+[AsF6]? The low-temperature fluorination of cyanuric chloride, (ClCN)3, with F2/AsF5 in SO2F2 solution yielded the salt [(ClCN)3F]+ [AsF6]? ( 1 ) essentially in quantitative yield. Compound 1 was identified by a low-temperature single crystal X-ray structure determination: R 3 c, trigonal, a = b = 10.4246(23) Å, c = 15.1850(24) Å, V = 1429.1(4) Å 3, Z = 6, RF = 0.056, Rw = 0.076 (for significant reflections), RF = 0.088, Rw = 0.079 (for all reflections). Fluorination of neat (ClCN)3 with [NF4]+ [Sb2F11]? yielded NF3, CClF3, SbF3, N2 and traces of CF4. A qualitative scale for the oxidizing strength of the oxidative fluorinators NF4+ and (XCN)3F+ (X = H, F, Cl) has been computed ab initio.  相似文献   

7.
The title compounds, poly­[[[bis(2‐methoxy­ethyl) ether]­lithium(I)]‐di‐μ3‐tri­fluoro­methanesulfonato‐lithium(I)], [Li2(CF3SO3)2(C6H14O3)]n, and poly­[[[bis(2‐methoxy­ethyl) ether]­lithium(I)]‐di‐μ3‐tri­fluoro­acetato‐dilithium(I)‐μ3‐tri­fluoro­acetato], [Li3(C2F3O2)3(C6H14O3)]n, consist of one‐dimensional polymer chains. Both structures contain five‐coordinate Li+ cations coordinated by a tridentate diglyme [bis(2‐methoxy­ethyl) ether] mol­ecule and two O atoms, each from separate anions. In both structures, the [Li(diglyme)X2]? (X is CF3SO3 or CF3CO2) fragments are further connected by other Li+ cations and anions, creating one‐dimensional chains. These connecting Li+ cations are coordinated by four separate anions in both compounds. The CF3SO3? and CF3CO2? anions, however, adopt different forms of cation coordination, resulting in differences in the connectivity of the structures and solvate stoichiometries.  相似文献   

8.
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   

9.
Aromatic ketones are enantioseletively hydrogenated in alcohols containing [RuX{(S,S)‐Tsdpen}(η6p‐cymene)] (Tsdpen=TsNCH(C6H5)CH(C6H5)NH2; X=TfO, Cl) as precatalysts. The corresponding Ru hydride (X=H) acts as a reducing species. The solution structures and complete spectral assignment of these complexes have been determined using 2D NMR (1H‐1H DQF‐COSY, 1H‐13C HMQC, 1H‐15N HSQC, and 1H‐19F HOESY). Depending on the nature of the solvents and conditions, the precatalysts exist as a covalently bound complex, tight ion pair of [Ru+(Tsdpen)(cymene)] and X?, solvent‐separated ion pair, or discrete free ions. Solvent effects on the NH2 chemical shifts of the Ru complexes and the hydrodynamic radius and volume of the Ru+ and TfO? ions elucidate the process of precatalyst activation for hydrogenation. Most notably, the Ru triflate possessing a high ionizability, substantiated by cyclic voltammetry, exists in alcoholic solvents largely as a solvent‐separated ion pair and/or free ions. Accordingly, its diffusion‐derived data in CD3OD reflect the independent motion of [Ru+(Tsdpen)(cymene)] and TfO?. In CDCl3, the complex largely retains the covalent structure showing similar diffusion data for the cation and anion. The Ru triflate and chloride show similar but distinct solution behavior in various solvents. Conductivity measurements and catalytic behavior demonstrate that both complexes ionize in CH3OH to generate a common [Ru+(Tsdpen)(cymene)] and X?, although the extent is significantly greater for X=TfO?. The activation of [RuX(Tsdpen)(cymene)] during catalytic hydrogenation in alcoholic solvent occurs by simple ionization to generate [Ru+(Tsdpen)(cymene)]. The catalytic activity is thus significantly influenced by the reaction conditions.  相似文献   

10.
Monocationic bis‐allyl complexes [Ln(η3‐C3H5)2(thf)3]+[B(C6X5)4]? (Ln=Y, La, Nd; X=H, F) and dicationic mono‐allyl complexes of yttrium and the early lanthanides [Ln(η3‐C3H5)(thf)6]2+[BPh4]2? (Ln=La, Nd) were prepared by protonolysis of the tris‐allyl complexes [Ln(η3‐C3H5)3(diox)] (Ln=Y, La, Ce, Pr, Nd, Sm; diox=1,4‐dioxane) isolated as a 1,4‐dioxane‐bridged dimer (Ln=Ce) or THF adducts [Ln(η3‐C3H5)3(thf)2] (Ln=Ce, Pr). Allyl abstraction from the neutral tris‐allyl complex by a Lewis acid, ER3 (Al(CH2SiMe3)3, BPh3) gave the ion pair [Ln(η3‐C3H5)2(thf)3]+[ER31‐CH2CH?CH2)]? (Ln=Y, La; ER3=Al(CH2SiMe3)3, BPh3). Benzophenone inserts into the La? Callyl bond of [La(η3‐C3H5)2(thf)3]+[BPh4]? to form the alkoxy complex [La{OCPh2(CH2CH?CH2)}2(thf)3]+[BPh4]?. The monocationic half‐sandwich complexes [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)(thf)2]+[B(C6X5)4]? (Ln=Y, La; X=H, F) were synthesized from the neutral precursors [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)2(thf)] by protonolysis. For 1,3‐butadiene polymerization catalysis, the yttrium‐based systems were more active than the corresponding lanthanum or neodymium homologues, giving polybutadiene with approximately 90 % 1,4‐cis stereoselectivity.  相似文献   

11.
Reactions of Trihalodithiatriazines and Trifluorotriazine with Halogen-Lewis-Acids . From the reaction of (XCN)(XSN)2 and MX5 (X = Cl, M = Sb; X = F, M = As) the corresponding salts [(XCN)(XSN)(SN)]+MX6? ( 2, 4 ) are obtained. (FCN)3 adds AsF5 to the nitrogen-bonded adduct (FCN)3 · AsF5 ( 6 ). The X-ray-structure of 6 is reported.  相似文献   

12.
A one step synthesis of [TcO3]+[SO3F]? is reported. The compound is volatile at room temperature and has according to calculations a tetrahedral coordination around Tc and a monodentate SO3F group. In the solid state the [SO3F]? anion bridges three [TcO3]+ cations, and vice versa. Space group P21/c, Z = 4, lattice dimensions at 173 K: a = 695.4(11), b = 808.6(12), c = 893.3(14) pm, β = 97.36(8)°.  相似文献   

13.
The synthesis of AsF3(SO3F)2 by the reaction AsF3 + S2O6F2→AsF3(SO3F)2 is described. Various alternate routes leading to similar arsenic (V) fluoride-fluorosulfates are discussed. All materials are clear, viscous, strongly associated liquids of the general formula AsFn(SO3F)5?n, with n ranging from about 2 to 4. The presence of fluorosulfate bridges is ascertained by IR and Raman spectra.The spectroscopic investigation is also extended to arsenic (III) fluoride- fluorosulfates.  相似文献   

14.
A series of cationic and neutral RuII complexes of the general formula [Ru(L)(X) (tBuCN)4]+X? and [Ru(L)(X)2(tBuCN)3)], that is, [Ru(CF3SO3){NCC(CH3)3}4(IMesH2)]+[CF3SO3]? ( 1 ), [Ru(CF3SO3){NCC(CH3)3}4(IMes)]+[CF3SO3]? ( 2 ), [RuCl{NCC(CH3)3}4(IMes)]+Cl? ( 3 ), [RuCl{NCC(CH3)3}4(IMesH2)+Cl?]/[RuCl2{NCC(CH3)3}3(IMesH2)] ( 4 ), and [Ru(NCO)2{NCC(CH3)3}3(IMesH2)] ( 5 ) (IMes=1,3‐dimesitylimidazol‐2‐ylidene, IMesH2=1,3‐dimesityl‐imidazolin‐2‐ylidene) have been synthesized and used as UV‐triggered precatalysts for the ring‐opening metathesis polymerization (ROMP) of different norborn‐2‐ene‐ and cis‐cyclooctene‐based monomers. The absorption maxima of complexes 1 – 5 were in the range of 245–255 nm and thus perfectly fit the emission band of the 254 nm UV source that was used for activation. Only the cationic RuII‐complexes based on ligands capable of forming μ2‐complexes such as 1 and 2 were found to be truly photolatent in ROMP. In contrast, complexes 3 – 5 could be activated by UV light; however, they also showed a low but significant ROMP activity in the absence of UV light. As evidenced by 1H and 13C NMR spectroscopy, the structure of the polymers obtained with either 1 or 2 are similar to those found in the corresponding polymers prepared by the action of [Ru(CF3SO3)2(IMesH2)(CH‐2‐(2‐PrO)‐C6H4)], which strongly suggest the formation of Ru‐based Grubbs‐type initiators in the course of the UV‐based activation process. Precatalysts that have the IMesH2 ligand showed significantly enhanced reactivity as compared with those based on the IMes ligand, which is in accordance with reports on the superior reactivity of IMesH2‐based Grubbs‐type catalysts compared with IMes‐based systems.  相似文献   

15.
Dihalogen(Pentafluorophenyl)sulfonium(IV) Hexafluoroarsenate C6F5SX2+AsF6? (X = Cl, Br) and Crystal Structure of Di(pentafluorophenyl)sulfane (C6F5)2S The preparation and spectroscopic characterisation of the halogensulfonium salts C6F5SCl2+AsF6? and C6F5SBr2+AsF6? is reported. The new salts are much more stable than their trifluoromethyl derivatives. In addition the crystal structure of (C6F5)2S is reported. Space group P43212, Z = 4, 478 unique observed diffractometer data, Rint. = 0.07, lattice constants: a = 569.0(5) pm, c = 3785.8(22) pm, V = 1225 times; 10?30 m3.  相似文献   

16.
Pentafluorethyl Sulfurtrifluoride: Synthesis and Reactions By oxidation of (C2F5S?)2 ( 1 ) with AgF2 at 0°C a mixture of C2F5SF3 ( 2 ) and C2F5SF5 ( 3 ) besides C2F5S(O)F ( 4 ) is formed. With elemental fluorine only 3 is isolated, an intermediate in this reaction is (C2F5SF4?)2 ( 5 ). At ?40 to ?30°C the mixture of 2, 3 and 4 was reacted with TASF and AsF5, to give TAS+ C2F5SF4? ( 6 ), TAS+ C2F5S(O)F2? ( 7 ) and C2F5SF2+AsF6? ( 8 ), respectively. While 6 and 7 decompose rapidly in solution even at low temperatures, of thermally stable 8 the solid state structure was determined by x-ray diffraction.  相似文献   

17.
Using [Ga(C6H5F)2]+[Al(ORF)4]?( 1 ) (RF=C(CF3)3) as starting material, we isolated bis‐ and tris‐η6‐coordinated gallium(I) arene complex salts of p‐xylene (1,4‐Me2C6H4), hexamethylbenzene (C6Me6), diphenylethane (PhC2H4Ph), and m‐terphenyl (1,3‐Ph2C6H4): [Ga(1,4‐Me2C6H4)2.5]+ ( 2+ ), [Ga(C6Me6)2]+ ( 3+ ), [Ga(PhC2H4Ph)]+ ( 4+ ) and [(C6H5F)Ga(μ‐1,3‐Ph2C6H4)2Ga(C6H5F)]2+ ( 52+ ). 4+ is the first structurally characterized ansa‐like bent sandwich chelate of univalent gallium and 52+ the first binuclear gallium(I) complex without a Ga?Ga bond. Beyond confirming the structural findings by multinuclear NMR spectroscopic investigations and density functional calculations (RI‐BP86/SV(P) level), [Ga(PhC2H4Ph)]+[Al(ORF)4]?( 4 ) and [(C6H5F)Ga(μ‐1,3‐Ph2C6H4)2Ga(C6H5F)]2+{[Al(ORF)4] ?}2 ( 5 ), featuring ansa‐arene ligands, were tested as catalysts for the synthesis of highly reactive polyisobutylene (HR‐PIB). In comparison to the recently published 1 and the [Ga(1,3,5‐Me3C6H3)2]+[Al(ORF)4]? salt ( 6 ) (1,3,5‐Me3C6H3=mesitylene), 4 and 5 gave slightly reduced reactivities. This allowed for favorably increased polymerization temperatures of up to +15 °C, while yielding HR‐PIB with high contents of terminal olefinic double bonds (α‐contents=84–93 %), low molecular weights (Mn=1000–3000 g mol?1) and good monomer conversions (up to 83 % in two hours). While the chelate complexes delivered more favorable results than 1 and 6 , the reaction kinetics resembled and thus concurred with the recently proposed coordinative polymerization mechanism.  相似文献   

18.
In the title compound, [(CH3)2(C7H7)NH][(C6F5)3B(OH)] or C9H14N+·C18HBF15O?, the distorted tetrahedral borate anions are strongly hydrogen bonded to the substituted ammonium cations. The N?O separation in the N—H?O hydrogen bond is 2.728 (3) Å.  相似文献   

19.
The methylamino diazonium cations [CH3N(H)N2]+ and [CF3N(H)N2]+ were prepared as their low‐temperature stable [AsF6]? salts by protonation of azidomethane and azidotrifluoromethane in superacidic systems. They were characterized by NMR and Raman spectroscopy. Unequivocal proof of the protonation site was obtained by the crystal structures of both salts, confirming the formation of alkylamino diazonium ions. The Lewis adducts CH3N3?AsF5 and CF3N3?AsF5 were also prepared and characterized by low‐temperature NMR and Raman spectroscopy, and also by X‐ray structure determination for CH3N3?AsF5. Electronic structure calculations were performed to provide additional insights. Attempted electrophilic amination of aromatics such as benzene and toluene with methyl‐ and trifluoromethylamino diazonium ions were unsuccessful.  相似文献   

20.
About Chemistry and Structure of Olefin-monocyano-dicarbonyl-ferrate Anions By the reactions of olenFe(CO)3 [olen = C5H8(isoprene), C7H10(cycloheptadiene-1,3), C8H14(2,5-dimethylhexadiene-1,3)] with sodium bis [trimethylsilyl]amide the new anions [olenFe(CO)2CN]? are formed. All so far known [olenFe(CO)2CN]? complexes [olen = C5H8(isoprene), C7H10[cycloheptadiene-1,3], C4H6(butadiene), C5H8(pentadiene-1,3), C6H8(cyclohexadine-1,3), C6H10(2,3-dimethylbutadiene), C8H8(cyclooctatetraene)] have fluctional structures in solution as shown by 13C NMR spectroscopic investigations. At low temperatures only the isomer exists, in which the CN? ligand and one of the two CO molecules occuppy the basal positions of a square pyramide together with 2 C atoms of the diene part.  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