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 406 毫秒
1.
Upon reaction of gaseous Me3SiF with the in situ prepared Lewis acid Al(ORF)3, the stable ion‐like silylium compound Me3Si‐F‐Al(ORF)3 1 forms. The Janus‐headed 1 is a readily available smart Lewis acid that differentiates between hard and soft nucleophiles, but also polymerizes isobutene effectively. Thus, in reactions of 1 with soft nucleophiles (Nu), such as phosphanes, the silylium side interacts in an orbital‐controlled manner, with formation of [Me3Si?Nu]+ and the weakly coordinating [F?Al(ORF)3] or [(FRO)3Al‐F‐Al(ORF)3] anions. If exchanged for hard nucleophiles, such as primary alcohols, the aluminum side reacts in a charge‐controlled manner, with release of FSiMe3 gas and formation of the adduct R(H)O?Al(ORF)3. Compound 1 very effectively initiates polymerization of 8 to 21 mL of liquid C4H8 in 50 mL of CH2Cl2 already at temperatures between ?57 and ?30 °C with initiator loads as low as 10 mg in a few seconds with 100 % yield but broad polydispersities.  相似文献   

2.
Phosphorane Iminato-Trichloroselenates(II): Syntheses and Crystal Structures of [SeCl(NPPh3)2]+SeCl3? and [Me3SiN(H)PMe3]2+[Se2Cl6]2? [SeCl(NPPh3)2]+SeCl3? has been synthesized by the reaction of Se2Cl2 with Me3SiNPPh3 in acetonitrile solution, forming orangered crystals, whereas red crystals of [Me3SiN(H)PMe3]2+[Se2Cl6]2? were obtained by the reaction of Me3SiNPMe3 with SeOCl2 in acetonitrile solution. Both complexes were characterized by X-ray structure determinations. [SeCl(NPPh3)2]+SeCl3?: Space group P21/n, Z = 4, structure solution with 7 489 observed unique reflections, R = 0.057. Lattice dimensions at ?60°C: a = 1 117.0; b = 2 241, c = 1 407.5 pm, β = 95.61°. In the cation [SeCl(NPPh3)2]+ the selenium atom is φ-tetrahedrally coordinated by the chlorine atom and by the nitrogen atoms of the phosphorane iminato ligands, whereas the anion SeCl3? has a T-shaped structure with φ-trigonal-bipyramidale surrounding of the selenium atom. [Me3SiN(H)PMe3]2+[Se2Cl6]2?: Space group P21/c, Z = 4, structure solution with 2 093 observed unique reflections, R = 0.080. Lattice dimensions at ?70°C: a = 956, b = 828, c = 1 973 pm, β = 93.80°. The structure consists of [Me3SiN(H)PMe3]+ ions and planar [Se2Cl6]2? anions, in which the selenium atoms are bridged nearly symmetrically by two chlorine atoms.  相似文献   

3.
The pnictocenium salts [Cp*PCl]+[μCl]? ( 1 a ), [Cp*PCl]+[ClAl(ORF)3]? ( 1 b ), [Cp*AsCl]+[ClAl(ORF)3]? ( 2 ), and [(Cp*)2P]+[μCl]? ( 3 ), in which Cp*=Me5C5, μCl=(FRO)3Al? Cl? Al(ORF)3, and ORF=OC(CF3)3, were prepared by halide abstraction from the respective halopnictines with the Lewis superacid PhF→Al(ORF)3. 1 The X‐ray crystal structures of 1 a , 2 , and 3 established that in the half as well as in the sandwich cations the Cp* rings are attached in an η2‐fashion. By using one or two equivalents of the Lewis acid, the two new weakly coordinating anions [μCl]? and [ClAl(ORF)3]? resulted. They also stabilize the highly reactive cations in PhF or 1,2‐F2C6H4 solution at room temperature. The chloride ion affinities (CIAs) of a range of classical strong Lewis acids were also investigated. The calculations are based on a set of isodesmic BP86/SV(P) reactions and a non‐isodesmic reference reaction assessed at the G3MP2 level.  相似文献   

4.
Attempts to prepare previously unknown simple and very Lewis acidic [RZn]+[Al(ORF)4]? salts from ZnR2, AlR3, and HO?RF delivered the ion‐like RZn(Al(ORF)4) (R=Me, Et; RF=C(CF3)3) with a coordinated counterion, but never the ionic compound. Increasing the steric bulk in RZn+ to R=CH2CMe3, CH2SiMe3, or Cp*, thus attempting to induce ionization, failed and led only to reaction mixtures including anion decomposition. However, ionization of the ion‐like EtZn(Al(ORF)4) compound with arenes yielded the [EtZn(arene)2]+[Al(ORF)4]? salts with arene=toluene, mesitylene, or o‐difluorobenzene (o‐DFB)/toluene. In contrast to the ion‐like EtZn(η3‐C6H6)(CHB11Cl11), which co‐crystallizes with one benzene molecule, the less coordinating nature of the [Al(ORF)4]? anion allowed the ionization and preparation of the purely organometallic [EtZn(arene)2]+ cation. These stable materials have further applications as, for example, initiators of isobutene polymerization. DFT calculations to compare the Lewis acidities of the zinc cations to those of a large number of organometallic cations were performed on the basis of fluoride ion affinity. The complexation energetics of EtZn+ with arenes and THF was assessed and related to the experiments.  相似文献   

5.
We report herein the synthesis and full characterization of the donor‐free Lewis superacids Al(ORF)3 with ORF=OC(CF3)3 ( 1 ) and OC(C5F10)C6F5 ( 2 ), the stabilization of 1 as adducts with the very weak Lewis bases PhF, 1,2‐F2C6H4, and SO2, as well as the internal C? F activation pathway of 1 leading to Al2(F)(ORF)5 ( 4 ) and trimeric [FAl(ORF)2]3 ( 5 , ORF=OC(CF3)3). Insights have been gained from NMR studies, single‐crystal structure determinations, and DFT calculations. The usefulness of these Lewis acids for halide abstractions has been demonstrated by reactions with trityl chloride (NMR; crystal structures). The trityl salts allow the introduction of new, heteroleptic weakly coordinating [Cl‐Al(ORF)3]? anions, for example, by hydride or alkyl abstraction reactions.  相似文献   

6.
By utilizing reaction mixtures, such as Me3Si–X/[Me3Si–X–SiMe3]+ (X=CN, OCN, SCN, and NNN), it was possible to prepare the first examples of bissilylated pseudohalonium cations in high yields. The structure and bonding of a whole series of salts containing pseudohalonium cations is discussed on the basis of experimentally observed (X‐ray diffraction, Raman, and IR spectroscopy, and mass spectrometry) and theoretically obtained data. Salts containing pseudohalonium cations are only stable in the presence of weakly coordinating anions, such as the well‐known tetrakis(pentafluorophenyl)borate, [B(C6F5)4]?.  相似文献   

7.
In a new oxidative route, Ag+[Al(ORF)4]? (RF=C(CF3)3) and metallic indium were sonicated in aromatic solvents, such as fluorobenzene (PhF), to give a precipitate of silver metal and highly soluble [In(PhF)n]+ salts (n=2, 3) with the weakly coordinating [Al(ORF)4]? anion in quantitative yield. The In+ salt and the known analogous Ga+[Al(ORF)4]? were used to synthesize a series of homoleptic PR3 phosphane complexes [M(PR3)n]+, that is, the weakly PPh3‐bridged [(Ph3P)3In–(PPh3)–In(PPh3)3]2+ that essentially contains two independent [In(PPh3)3]+ cations or, with increasing bulk of the phosphane, the carbene‐analogous [M(PtBu3)2]+ (M=Ga, In) cations. The MI? P distances are 27 to 29 pm longer for indium, and thus considerably longer than the difference between their tabulated radii (18 pm). The structure, formation, and frontier orbitals of these complexes were investigated by calculations at the BP86/SV(P), B3LYP/def2‐TZVPP, MP2/def2‐TZVPP, and SCS‐MP2/def2‐TZVPP levels.  相似文献   

8.
The straightforward synthesis of the cationic, purely organometallic NiI salt [Ni(cod)2]+[Al(ORF)4] was realized through a reaction between [Ni(cod)2] and Ag[Al(ORF)4] (cod=1,5‐cyclooctadiene). Crystal‐structure analysis and EPR, XANES, and cyclic voltammetry studies confirmed the presence of a homoleptic NiI olefin complex. Weak interactions between the metal center, the ligands, and the anion provide a good starting material for further cationic NiI complexes.  相似文献   

9.
The long-postulated reactive intermediates of polar fluoride-inititated trifluoromethylations with Me3SiCF3 (see scheme) were identified by NMR spectroscopy as [Me3Si(CF3)F] and [Me3Si(CF3)2].  相似文献   

10.
Metal Ampoules as Mini‐Autoclaves: Syntheses and Crystal Structures of [Al(NH3)4Cl2][Al(NH3)2Cl4] and (NH4)2[Al(NH3)4Cl2][Al(NH3)2Cl4]Cl2 The salts [Al(NH3)4Cl2]+[Al(NH3)2Cl4]≡AlCl3 · 3 NH3 ( 1 ) and (NH4+)2[Al(NH3)4Cl2]+[Al(NH3)2Cl4](Cl)2≡ AlCl3 · 3 NH3 · (NH4)Cl ( 2 ) have been obtained as single crystals during the reactions of aluminum and aluminum trichloride, respectively, with ammonium chloride in sealed Monel metal containers. The crystal structure of 1 was determined again [triclinic, P‐1; a = 574.16(10); b = 655.67(12); c = 954.80(16) pm; α = 86.41(2); β = 87.16(2); γ = 84.89(2)°], that of 2 for the first time [monoclinic, I2/m; a = 657.74(12); b = 1103.01(14); c = 1358.1(3) pm; β = 103.24(2)°].  相似文献   

11.
Schnöckel's [(AlCp*)4] and Jutzi's [SiCp*][B(C6F5)4] (Cp*=C5Me5) are landmarks in modern main-group chemistry with diverse applications in synthesis and catalysis. Despite the isoelectronic relationship between the AlCp* and the [SiCp*]+ fragments, their mutual reactivity is hitherto unknown. Here, we report on their reaction giving the complex salts [Cp*Si(AlCp*)3][WCA] ([WCA]=[Al(ORF)4] and [F{Al(ORF)3}2]; RF=C(CF3)3). The tetrahedral [SiAl3]+ core not only represents a rare example of a low-valent silicon-doped aluminium-cluster, but also—due to its facile accessibility and high stability—provides a convenient preparative entry towards low-valent Si−Al clusters in general. For example, an elusive binuclear [Si2(AlCp*)5]2+ with extremely short Al−Si bonds and a high negative partial charge at the Si atoms was structurally characterised and its bonding situation analysed by DFT. Crystals of the isostructural [Ge2(AlCp*)5]2+ dication were also obtained and represent the first mixed Al−Ge cluster.  相似文献   

12.
Bis(dimethylamino)trifluoro sulfonium Salts: [CF3S(NMe2)2]+[Me3SiF2], [CF3S(NMe2)2]+ [HF2] and [CF3S(NMe2)2]+[CF3S] From the reaction of CF3SF3 with an excess of Me2NSiMe3 [CF3(NMe2)2]+[Me3SiF2] (CF3‐BAS‐fluoride) ( 5 ), from CF3SF3/CF3SSCF3 and Me2NSiMe3 [CF3S(NMe2)2]+‐ [CF3S] ( 7 ) are isolated. Thermal decomposition of 5 gives [CF3S(NMe2)2]+ [HF2] ( 6 ). Reaction pathways are discussed, the structures of 5 ‐ 7 are reported.  相似文献   

13.
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   

14.
The trapping of a silicon(I) radical with N‐heterocyclic carbenes is described. The reaction of the cyclic (alkyl)(amino) carbene [cAACMe] (cAACMe=:C(CMe2)2(CH2)NAr, Ar=2,6‐i Pr2C6H3) with H2SiI2 in a 3:1 molar ratio in DME afforded a mixture of the separated ion pair [(cAACMe)2Si:.]+I ( 1 ), which features a cationic cAAC–silicon(I) radical, and [cAACMe−H]+I. In addition, the reaction of the NHC–iodosilicon(I) dimer [IAr(I)Si:]2 (IAr=:C{N(Ar)CH}2) with 4 equiv of IMe (:C{N(Me)CMe}2), which proceeded through the formation of a silicon(I) radical intermediate, afforded [(IMe)2SiH]+I ( 2 ) comprising the first NHC–parent‐silyliumylidene cation. Its further reaction with fluorobenzene afforded the CAr−H bond activation product [1‐F‐2‐IMe‐C6H4]+I ( 3 ). The isolation of 2 and 3 confirmed the reaction mechanism for the formation of 1 . Compounds 1 – 3 were analyzed by EPR and NMR spectroscopy, DFT calculations, and X‐ray crystallography.  相似文献   

15.
Ion‐like ethylzinc(II) compounds with weakly coordinating aluminates [Al(ORF)4]? and [(RFO)3Al‐F‐Al(ORF)3]? (RF=C(CF3)3) were synthesized in a one‐pot reaction and fully characterized by single‐crystal X‐ray diffraction, NMR and vibrational spectroscopy, and by quantum chemical calculations. The catalytic activity of ion‐like Et‐Zn[Al(ORF)4] in intermolecular hydroamination and in the unusual double hydroamination of anilines and alkynes was investigated. Favorable performance was also found in comparison to the Et2Zn/ [PhNMe2H]+[B(C6F5)4]? system generated in situ at lower catalyst loadings of 2.5 mol %.  相似文献   

16.
Compounds including the free or coordinated gas‐phase cations [Ag(η2‐C2H4)n]+ (n=1–3) were stabilized with very weakly coordinating anions [A]? (A=Al{OC(CH3)(CF3)2}4, n=1 ( 1 ); Al{OC(H)(CF3)2}4, n=2 ( 3 ); Al{OC(CF3)3}4, n=3 ( 5 ); {(F3C)3CO}3Al‐F‐Al{OC(CF3)3}3, n=3 ( 6 )). They were prepared by reaction of the respective silver(I) salts with stoichiometric amounts of ethene in CH2Cl2 solution. As a reference we also prepared the isobutene complex [(Me2C?CH2)Ag(Al{OC(CH3)(CF3)2}4)] ( 2 ). The compounds were characterized by multinuclear solution‐NMR, solid‐state MAS‐NMR, IR and Raman spectroscopy as well as by their single crystal X‐ray structures. MAS‐NMR spectroscopy shows that the [Ag(η2‐C2H4)3]+ cation in its [Al{OC(CF3)3}4]? salt exhibits time‐averaged D3h‐symmetry and freely rotates around its principal z‐axis in the solid state. All routine X‐ray structures (2θmax.<55°) converged within the 3σ limit at C?C double bond lengths that were shorter or similar to that of free ethene. In contrast, the respective Raman active C?C stretching modes indicated red‐shifts of 38 to 45 cm?1, suggesting a slight C?C bond elongation. This mismatch is owed to residual librational motion at 100 K, the temperature of the data collection, as well as the lack of high angular data owing to the anisotropic electron distribution in the ethene molecule. Therefore, a method for the extraction of the C?C distance in [M(C2H4)] complexes from experimental Raman data was developed and meaningful C?C distances were obtained. These spectroscopic C?C distances compare well to newly collected X‐ray data obtained at high resolution (2θmax.=100°) and low temperature (100 K). To complement the experimental data as well as to obtain further insight into bond formation, the complexes with up to three ligands were studied theoretically. The calculations were performed with DFT (BP86/TZVPP, PBE0/TZVPP), MP2/TZVPP and partly CCSD(T)/AUG‐cc‐pVTZ methods. In most cases several isomers were considered. Additionally, [M(C2H4)3] (M=Cu+, Ag+, Au+, Ni0, Pd0, Pt0, Na+) were investigated with AIM theory to substantiate the preference for a planar conformation and to estimate the importance of σ donation and π back donation. Comparing the group 10 and 11 analogues, we find that the lack of π back bonding in the group 11 cations is almost compensated by increased σ donation.  相似文献   

17.
The complexes Ag(L)n[WCA] (L=P4S3, P4Se3, As4S3, and As4S4; [WCA]=[Al(ORF)4] and [F{Al(ORF)3}2]; RF=C(CF3)3; WCA=weakly coordinating anion) were tested for their performance as ligand-transfer reagents to transfer the poorly soluble nortricyclane cages P4S3, P4Se3, and As4S3 as well as realgar As4S4 to different transition-metal fragments. As4S4 and As4S3 with the poorest solubility did not yield complexes. However, the more soluble silver-coordinated P4S3 and P4Se3 cages were transferred to the electron-poor Fp+ moiety ([CpFe(CO)2]+). Thus, reaction of the silver salt in the presence of the ligand with Fp−Br yielded [Fp−P4S3][Al(ORF)4] ( 1 a ), [Fp−P4S3][F(Al(ORF)3)2] ( 1 b ), and [Fp−P4Se3][Al(ORF)4] ( 2 ). Reactions with P4S3 also yielded [FpPPh3−P4S3][Al(ORF)4] ( 3 ), a complex with the more electron-rich monophosphine-substituted Fp+ analogue [FpPPh3]+ ([CpFe(PPh3)(CO)]+). All complex salts were characterized by single-crystal XRD, NMR, Raman, and IR spectroscopy. Interestingly, they show characteristic blueshifts of the vibrational modes of the cage, as well as structural contractions of the cages upon coordination to the Fp/FpPPh3 moieties, which oppose the typically observed cage expansions that lead to redshifts in the spectra. Structure, bonding, and thermodynamics were investigated by DFT calculations, which support the observed cage contractions. Its reason is assigned to σ and π donation from the slightly P−P and P−E antibonding P4E3-cage HOMO (e symmetry) to the metal acceptor fragment.  相似文献   

18.
Investigations on the Reactivity of [Me2AlP(SiMe3)2]2 with Base‐stabilized Organogalliumhalides and ‐hydrides [Me2AlP(SiMe3)2]2 ( 1 ) reacts with dmap?Ga(Cl)Me2, dmap?Ga(Me)Cl2, dmap?GaCl3 and dmap?Ga(H)Me2 with Al‐P bond cleavage and subsequent formation of heterocyclic [Me2GaP(SiMe3)2]2 ( 2 ) as well as dmap?AlMexCl3?x (x = 3 8 ; 2 3 ; 1 4 ; 0 5 ). The reaction between equimolar amounts of dmap?Al(Me2)P(SiMe3)2 and dmap?Ga(t‐Bu2)Cl yield dmap?Ga(t‐Bu2)P(SiMe3)2 ( 6 ) and dmap?AlMe2Cl ( 3 ). 2 – 8 were characterized by NMR spectroscopy, 2 and 6 also by single crystal X‐ray diffraction.  相似文献   

19.
Syntheses and Crystal Structures of [μ‐(Me3SiCH2Sb)5–Sb1,Sb3–{W(CO)5}2] and [{(Me3Si)2CHSb}3Fe(CO)4] – Two Cyclic Complexes with Antimony Ligands cyclo‐(Me3SiCH2Sb)5 reacts with [(THF)W(CO)5] (THF = tetrahydrofuran) to form cyclo‐[μ‐(Me3SiCH2Sb)5–Sb1,Sb3–{W(CO)5}2] ( 1 ). The heterocycle cyclo‐ [{(Me3Si)2CHSb}3Fe(CO)4] ( 2 ) is formed by an insertion reaction of cyclo‐[(Me3Si)2CHSb]3 and [Fe2(CO)9]. The crystal structures of 1 and 2 are reported.  相似文献   

20.
Coordination Chemistry of P‐rich Phosphanes and Silylphosphanes. XXII. The Formation of [η2‐{tBu–P=P–SiMe3}Pt(PR3)2] from (Me3Si)tBuP–P=P(Me)tBu2 and [η2‐{C2H4}Pt(PR3)2] (Me3Si)tBuP–P = P(Me)tBu2 reacts with [η2‐{C2H4}Pt(PR3)2] yielding [η2‐{tBu–P=P–SiMe3}Pt(PR3)2]. However, there is no indication for an isomer which would be the analogue to the well known [η2‐{tBu2P–P}Pt(PPh3)2]. The syntheses and NMR data of [η2‐{tBu–P=P–SiMe3}Pt(PPh3)2] and [η2‐{tBu–P=P–SiMe3}Pt(PMe3)2] as well as the results of the single crystal structure determination of [η2‐{tBu–P=P–SiMe3}Pt(PPh3)2] are reported.  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