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 599 毫秒
1.
Two very rare cases of barium boryloxides, the homoleptic [Ba(OB{CH(SiMe3)2}2)2⋅C7H8] and the heteroleptic [{LONO4}BaOB{CH(SiMe3)2}2] stabilised by the multidentate aminoetherphenolate {LONO4}, are presented, and their structural properties are discussed. The electron-deficient [Ba(OB{CH(SiMe3)2}2)2⋅C7H8] shows, in particular, resilient η6-coordination of the toluene molecule. Together with its amido parents [Ba{N(SiMe3)2}2⋅thf2] and [Ba{N(SiMe3)2}2]2, this complex catalyses the fast and chemoselective dehydrocoupling of borinic acids R2BOH and hydrosilanes HSiR′3, yielding borasiloxanes R2BOSiR′3 in a controlled fashion. The assessment of substrate scope indicates that, for now, the reaction is limited to bulky borinic acids. Kinetic analysis shows that the rate-limiting step of the catalytic manifold traverses a dinuclear transition state. A detailed mechanistic scenario is proposed on the basis of DFT computations, the results of which are fully consistent with experimental data. It consists of a stepwise process with rate-determining nucleophilic attack of a metal-bound O-atom onto the incoming hydrosilane, involving throughout dinuclear catalytically active species.  相似文献   

2.
Barium complexes ligated by bulky boryloxides [OBR2] (where R=CH(SiMe3)2, 2,4,6-iPr3-C6H2 or 2,4,6-(CF3)3-C6H2), siloxide [OSi(SiMe3)3], and/or phenoxide [O-2,6-Ph2-C6H3], have been prepared. A diversity of coordination patterns is observed in the solid state for both homoleptic and heteroleptic complexes, with coordination numbers ranging between 2 and 4. The identity of the bridging ligand in heteroleptic dimers [Ba(μ2-X1)(X2)]2 depends largely on the given pair of ligands X1 and X2. Experimentally, the propensity to fill the bridging position increases according to [OB{CH(SiMe3)2}2)]<[N(SiMe3)2]<[OSi(SiMe3)3]<[O(2,6-Ph2-C6H3)]<[OB(2,4,6-iPr3-C6H2)2]. This trend is the overall expression of 3 properties: steric constraints, electronic density and σ- and π-donating capability of the negatively charged atom, and ability to generate Ba ⋅ ⋅ ⋅ F, Ba ⋅ ⋅ ⋅ C(π) or Ba ⋅ ⋅ ⋅ H−C secondary interactions. The comparison of the structural motifs in the complexes [Ae{μ2-N(SiMe3)2}(OB{CH(SiMe3)2}2)]2 (Ae = Mg, Ca, Sr and Ba) suggest that these observations may be extended to all alkaline earths. DFT calculations highlight the largely prevailing ionic character of ligand-Ae bonding in all compounds. The ionic character of the Ae-ligand bond encourages bridging coordination, whereas the number of bridging ligands is controlled by steric factors. DFT computations also indicate that in [Ba(μ2-X1)(X2)]2 heteroleptic dimers, ligand predilection for bridging vs. terminal positions is dictated by the ability to establish secondary interactions between the metals and the ligands.  相似文献   

3.
Synthesis and Structure of Tetrameric Tris(trimethylsilyl)indium(I) and of New Silyl substituted Indium Compounds The reaction of InCp* with [LiSi(SiMe3)3·3thf] yielded in the first silylsubstituted tetrahedrane of indium [In4{Si(SiMe3)3}4] ( 1 ). It crystallizes together with [In{Si(SiMe3)3}3] ( 2 ) in dark green crystals. Colourless crystals of [Li(OH)(OSiMe3)In{Si(SiMe3)3}2]2 ( 3 ) were isolated as a byproduct from this reaction. It's structural core are three connected four membered rings made up of In‐, Li‐ and O‐atoms. From the reaction of [InOSO2CF3] with [LiSi(SiMe3)3·3thf] colourless crystals of [In{Si(SiMe3)3}2OSO2CF3·thf] ( 4 ) were isolated. InCp* reacted with [LiSiMe(SiMe3)2·3thf] to form the orange‐coloured monoindane [In{SiMe(SiMe3)2}3] ( 5 ). 1 – 4 were characterized by X‐ray crystal structure analyses.  相似文献   

4.
The generation and properties of the Cp2Zr{CH(SiMe3)2}+ cation are described. An X-ray crystallographic analysis shows that the carborane salt [Cp2Zr{CH(SiMe3)2}][HCB11Me5Br6] contains an agostic Zr-μ-Me-Si interaction in the solid state. Low temperature NMR spectra of the borate salt [Cp2Zr{CH(SiMe3)2}][B(C6F5)4] show that this interaction is retained in solution. Variable temperature NMR spectra establish that the SiMe2(μ-Me) and unbound SiMe3 units of Cp2Zr{CH(SiMe3)2}+ exchange by a “pivot” process involving partial rotation around the Zr-CH(SiMe3)2 bond, with a barrier of ΔG = 9.2(1) kcal/mol at −89 °C. Cp2Zr{CH(SiMe3)2}+ does not coordinate alkenes or alkynes.  相似文献   

5.
The first unsupported barium siloxide, the homoleptic dimer [Ba22-OSi(SiMe3)3}3{OSi(SiMe3)3}], is presented, and its structural features are discussed in the light of DFT computations. This complex, together with the related [Ba{μ2-OSi(SiMe3)3}{N(SiMe3)2}]2 and their parent [Ba{N(SiMe3)2}2]2, mediates the formation of asymmetric siloxanes R3Si−O−SiR′3 through the first case of main group metal-mediated dehydrocoupling of silanols and hydrosilanes. Early kinetic analysis highlights an unusual catalytic manifold.  相似文献   

6.
The catalysis of a silica‐supported chromium system {Cr[CH(SiMe3)2]3/SiO2} was compared with a silica‐supported chromium oxide catalyst, the Phillips catalyst (CrO3/SiO2). This catalyst was prepared by the calcining of the typical silica support used for the Phillips catalyst at 600 °C and by the support of tris[bis(trimethylsilyl)methyl]chromium(III) {Cr[CH(SiMe3)2]3} on the silica. In the slurry‐phase polymerization, this catalyst conducted the polymerization of ethylene at a high activity without organoaluminum compounds as cocatalysts or scavengers. The activity per Cr was about 6–7 times higher than that of the Phillips catalyst. Upon the introduction of hydrogen to the system, the molecular weight of polyethylene did not change with the Phillips catalyst, but it decreased with the Cr[CH(SiMe3)2]3/SiO2 catalyst. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 413–419, 2003  相似文献   

7.
First‐row two‐coordinate complexes are attracting much interest. Herein, we report the high‐yield isolation of the linear two‐coordinate iron(I) complex salt [K(L)][Fe{N(SiMe3)2}2] (L=18‐crown‐6 or crypt‐222) through the reduction of either [Fe{N(SiMe3)2}2] or its three‐coordinate phosphine adduct [Fe{N(SiMe3)2}2(PCy3)]. Detailed characterization is gained through X‐ray diffraction, variable‐temperature NMR spectroscopy, and magnetic susceptibility studies. One‐ and two‐electron oxidation through reaction with I2 is further found to afford the corresponding iodo iron(II) and diiodo iron(III) complexes.  相似文献   

8.
First‐row two‐coordinate complexes are attracting much interest. Herein, we report the high‐yield isolation of the linear two‐coordinate iron(I) complex salt [K(L)][Fe{N(SiMe3)2}2] (L=18‐crown‐6 or crypt‐222) through the reduction of either [Fe{N(SiMe3)2}2] or its three‐coordinate phosphine adduct [Fe{N(SiMe3)2}2(PCy3)]. Detailed characterization is gained through X‐ray diffraction, variable‐temperature NMR spectroscopy, and magnetic susceptibility studies. One‐ and two‐electron oxidation through reaction with I2 is further found to afford the corresponding iodo iron(II) and diiodo iron(III) complexes.  相似文献   

9.
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   

10.
Alkaline‐earth (Ae=Ca, Sr, Ba) complexes are shown to catalyse the chemoselective cross‐dehydrocoupling (CDC) of amines and hydrosilanes. Key trends were delineated in the benchmark couplings of Ph3SiH with pyrrolidine or tBuNH2. Ae{E(SiMe3)2}2 ? (THF)x (E=N, CH; x=2–3) are more efficient than {N^N}Ae{E(SiMe3)2} ? (THF)n (E=N, CH; n=1–2) complexes (where {N^N}?={ArN(o‐C6H4)C(H)=NAr}? with Ar=2,6‐iPr2‐C6H3) bearing an iminoanilide ligand, and alkyl precatalysts are better than amido analogues. Turnover frequencies (TOFs) increase in the order Ca<Sr<Ba. Ba{CH(SiMe3)2}2 ? (THF)3 displays the best performance (TOF up to 3600 h?1). The substrate scope (>30 products) includes diamines and di(hydrosilane)s. Kinetic analysis of the Ba‐promoted CDC of pyrrolidine and Ph3SiH shows that 1) the kinetic law is rate=k[Ba]1[amine]0[hydrosilane]1, 2) electron‐withdrawing p‐substituents on the arylhydrosilane improve the reaction rate and 3) a maximal kinetic isotopic effect (kSiH/kSiD=4.7) is seen for Ph3SiX (X=H, D). DFT calculations identified the prevailing mechanism; instead of an inaccessible σ‐bond‐breaking metathesis pathway, the CDC appears to follow a stepwise reaction path with N?Si bond‐forming nucleophilic attack of the catalytically competent Ba pyrrolide onto the incoming silane, followed by rate limiting hydrogen‐atom transfer to barium. The participation of a Ba silyl species is prevented energetically. The reactivity trend Ca<Sr<Ba results from greater accessibility of the metal centre and decreasing Ae?Namide bond strength upon descending Group 2.  相似文献   

11.
Treatment of AsP3 with 0.75 equivalents of [{GaC(SiMe3)3}4] resulted in selective insertion of three equivalents of {GaC(SiMe3)3} into the three As? P bonds to give [As{GaC(SiMe3)3}3P3] ( 1 ‐As) with an intact cyclo‐P3 ring. This yellow compound has been characterized by NMR spectroscopy, combustion analysis, single‐crystal X‐ray diffraction, UV/Vis spectroscopy, Raman spectroscopy, and cyclic voltammetry (THF, 0.2 M [TBA][B(C6F5)4]; TBA=tetrabutyl ammonium). Computational models of 1 ‐As and the isomeric [P{GaC(SiMe3)3}3AsP2] ( 1 ‐P) have been investigated as well, revealing several interesting electronic features of these cage molecules. Following from the cyclic voltammetry studies of 1 ‐As that highlight an irreversible two‐electron reduction at ?2.2 V versus Fc/Fc+, treatment with one equivalent of [Mg(C14H10)(thf)3] resulted in two‐electron reduction to provide [As{GaC(SiMe3)3}3P3Mg(thf)3] ( 2 ), in which the Mg2+ ion has inserted into one of the P? P bonds of the cyclo‐P3 ring. It was also found that treatment of AsP3 or P4 with one equivalent of [{GaC(SiMe3)3}4] resulted in formation of the quadruple insertion products [As{GaC(SiMe3)3}4P3] ( 3 ) and [P{GaC(SiMe3)3}4P3] ( 4 ), respectively.  相似文献   

12.
Trialkylhydridoalanates RxR′3?xAlH? [R = CMe3; R′ = CH(SiMe3)2] The very strong base tert-butyl lithium reacts in the presence of chelating tetramethylethylendiamine with the aluminium organyls Al[CH(SiMe3)2]2CMe3 1 and Al[CH(SiMe3)2](CMe3)2 2 not under proton abstraction from the C? H acidic elementorganic substituent, but under β-elimination and addition of the thereby formed LiH to the coordinatively unsaturated aluminium atom. Two alanates — [Hal{CH(SiMe3)2}2CMe3]? 3 and [HAl{CH(SiMe3)2}(CMe3)2]? 4 each with Li(TMEDA)2 as counterion — were isolated; they exhibit separate anions and cations in solid state as shown by a crystal structure determination on 3 . In absence of the chelating amine tert-butyl lithium decomposes under the catalytic effect of the aluminium compound to LiH, which does not add to aluminium and precipitates in a reactive form.  相似文献   

13.
The reaction of YbCl3 with two equivalents of NaN‐(SiMe3)2 has afforded a mixture of several ytterbium bis(trimethylsilyl) amides with the known complexes [Yb{N(SiMe3)2}2(μ‐Cl)(thf)]2 ( 1 ) and [Yb{N(SiMe3)2}3]( 4 ) as the main products and the cluster compound [Yb3Cl4O{N(SiMe3)2}3(thf)3]( 2 ) as a minor product. Treatment of 1 and 2 with hot n‐heptane gave the basefree complex [Yb{N(SiMe3)2}2(μ‐Cl)]2 ( 3 ) in small yield. The structures of compounds 1—4 and the related peroxo complex [Yb2{N(SiMe3)2}4(μ‐O2)(thf)2]( 5 ) have been investigated by single crystal X‐ray diffraction. In the solid‐state, 3 shows chlorobridged dimers with terminal amido ligands (av. Yb—Cl = 262.3 pm, av. Yb—N = 214.4 pm). Additional agostic interactions are observed from the ytterbium atoms to four methyl carbon atoms of the bis(trimethylsilyl)amido groups (Yb···C = 284—320 pm). DFT calculations have been performed on suitable model systems ([Yb2(NH2)4(μ‐Cl)2(OMe2)2]( 1m ), [Yb2(NH2)4(μ‐Cl)2]( 3m ), [Yb‐(NH2)3]( 4m ), [Yb2(NH24(μ‐O2)(OMe2)2]( 5m ), [Yb{N‐(SiMe3)2}2Cl] ( 3m/2 ) and Ln(NH2)2NHSiMe3 (Ln = Yb ( 6m ), Y ( 7m )) in order to rationalize the different experimentally observed Yb—N distances, to support the assignment of the O—O stretching vibration (775 cm ‐1) in the Raman spectrum of complex 5 and to examine the nature of the agostic‐type interactions in σ‐donorfree 3 .  相似文献   

14.
Reaction of ArN3 (Ar = Ph, p-MeC6H4, 1-naphthyl) with [Li{Si(SiMe3)3}(thf)3] yielded lithium amides [Li{N(Ar)Si(SiMe3)3}L] (L = tmeda or (thf)2). Similar treatment of o-phenylene diazide with 2 equiv. of [Li{Si(SiMe3)3}(thf)3] formed dilithium diamide complex 4. Reaction between o-Me3SiOC6H4N3 and [Li{Si(SiMe3)3}(thf)3] afforded, via 1,4-trimethylsilyl migration from oxygen to nitrogen, [Li{OC6H4{N(SiMe3)Si(SiMe3)3}-2}]2 (5). The structures of complexes 3 and 5 have been determined by single crystal X-ray diffraction techniques.  相似文献   

15.
The branched triphosphanyltetrasilane PhSi(SiMe2PH2)3 ( 1 ) could be obtained in a three‐stage synthesis. It was characterised by multi‐nuclear NMR spectroscopy, mass spectrometry and IR spectroscopy. Deprotonation of 1 with GaiPr3 or [M{N(SiMe3)2}2(thf)2] (M = Ca, Sr, Ba) yields new phosphorus bridged polynuclear complexes of these metals with phosphorus atoms connected through tetrasilane fragments. While trinuclear complexes with single deprotonated phosphanyl groups could be obtained from the reactions of 1 with GaiPr3, calcium or barium silazanide (compounds 2 , 3 and 5 ), the tetranuclear complex [Sr4{PhSi(SiMe2PH)2(SiMe2P)}2(dme)6] ( 4 ) was formed in the reaction of 1 with strontium silazanide. In this compound, two of six phosphorus atoms are deprotonated twice. Compounds 2 – 5 were characterised by single‐crystal X‐ray diffraction, elemental analysis as well as IR spectroscopy and as far as possible by NMR spectroscopic techniques.  相似文献   

16.
A set of calcium and barium complexes containing the fluoroarylamide N(C6F5)2 is presented. These compounds illustrate the key role of stabilising M⋅⋅⋅F−C secondary interactions in the construction of low-coordinate alkaline earth complexes. The nature of Ca⋅⋅⋅F−C bonding in calcium complexes is examined in the light of structural data, bond valence sum (BVS) analysis and DFT computations. The molecular structures of [Ca{N(C6F5)2}2(Et2O)2] ( 4 ′), [Ca{μ-N(SiMe3)2}{N(C6F5)2}]2 ( 52 ), [Ba{μ-N(C6F5)2}{N(C6F5)2}⋅toluene]2 ( 62 ), [{BDIDiPP}CaN(C6F5)2]2 ( 72 ), [{N^NDiPP}CaN(C6F5)2]2 ( 82 ), and [Ca{μ-OB(CH(SiMe3)2)2}{N(C6F5)2}]2 ( 92 ), where {BDIDiPP} and {N^NDiPP} are the bidentate ligands CH[C(CH3)NDipp]2 and DippNC6H4CNDipp (Dipp=2,6-iPr2-C6H3), are detailed. Complex 62 displays strong Ba⋅⋅⋅F−C contacts at around 2.85 Å. The calcium complexes feature also very short intramolecular Ca−F interatomic distances at around 2.50 Å. In addition, the three-coordinate complexes 72 and 82 form dinuclear structures due to intermolecular Ca⋅⋅⋅F−C contacts. BVS analysis shows that Ca⋅⋅⋅F−C interactions contribute to 15–20 % of the bonding pattern around calcium. Computations demonstrate that Ca⋅⋅⋅F−C bonding is mostly electrostatic, but also contains a non-negligible covalent contribution. They also suggest that Ca⋅⋅⋅F−C are the strongest amongst the range of weak Ca⋅⋅⋅X (X=F, H, Cπ) secondary interactions, due to the high positive charge of Ca2+ which favours electrostatic interactions.  相似文献   

17.
The characterization of the unstable NiII bis(silylamide) Ni{N(SiMe3)2}2 ( 1 ), its THF complex Ni{N(SiMe3)2}2(THF) ( 2 ), and the stable bis(pyridine) derivative trans‐Ni{N(SiMe3)2}2(py)2 ( 3 ), is described. Both 1 and 2 decompose at ca. 25 °C to a tetrameric NiI species, [Ni{N(SiMe3)2}]4 ( 4 ), also obtainable from LiN(SiMe3)2 and NiCl2(DME). Experimental and computational data indicate that the instability of 1 is likely due to ease of reduction of NiII to NiI and the stabilization of 4 through dispersion forces.  相似文献   

18.
Thermolysis of Sterically Stressed Alanates; Synthesis of Two New 1-Sila-3-alanata-cyclobutane Derivatives with Four-membered AlC2Si-Heterocycles The reaction of high shielded alkyl or aryl alanes with LiCH(SiMe3)2 in the presence of the chelating N,N′,N″-trimethyl-triazinane yields the sterically stressed alanates [(Me3C)2Al{CH(SiMe3)2}2]? 12 and [R? Al{CH(SiMe3)2}3]? (R = Me3SiCH2 13 , Et 14 , Me 15 , C6H5 16 ) each with a Li(triazinane)2 counter ion. On thermolysis of the sterically most shielded derivatives 12 and 13 at 130 to 150°C one equivalent of bis(trimethylsilyl)methane is liberated, and by deprotonation of methyl groups carbanionic species are formed, which are stabilized by intramolecular coordination to the unsaturated aluminium atoms under formation of AlC2Si heterocycles ( 19 and 20 ). 20 was characterized by a single crystal structure determination. The remaining alanates give under similar conditions either under dismutation the recently published heterocycle 1 with two intact CH(SiMe3)2 groups ( 14 and 15 ) or a methyl alanate by the replacement of a elementorganic substituent ( 16 ).  相似文献   

19.
[{N^N}M(X)(thf)n] alkyl (X=CH(SiMe3)2) and amide (X=N(SiMe3)2) complexes of alkaline earths (M=Ca, Sr, Ba) and divalent rare earths (YbII and EuII) bearing an iminoanilide ligand ({N^N}?) are presented. Remarkably, these complexes proved to be kinetically stable in solution. X‐ray diffraction studies allowed us to establish size–structure trends. Except for one case of oxidation with [{N^N}YbII{N(SiMe3)2}(thf)], all these complexes are stable under the catalytic conditions and constitute effective precatalysts for the cyclohydroamination of terminal aminoalkenes and the intermolecular hydroamination and intermolecular hydrophosphination of activated alkenes. Metals with equal sizes across alkaline earth and rare earth families display almost identical apparent catalytic activity and selectivity. Hydrocarbyl complexes are much better catalyst precursors than their amido analogues. In the case of cyclohydroamination, the apparent activity decreases with metal size: Ca>Sr>Ba, and the kinetic rate law agrees with RCHA=k[precatalyst]1[aminoalkene]1. The intermolecular hydroamination and hydrophosphination of styrene are anti‐Markovnikov regiospecific. In both cases, the apparent activity increases with the ionic radius (Ca<Sr<Ba) but the rate laws are different, and obey RHA=k[styrene]1[amine]1[precatalyst]1 and RHP=k[styrene]1[HPPh2]0[precatalyst]1, respectively. Mechanisms compatible with the rate laws and kinetic isotopic effects are proposed. [{N^N}Ba{N(SiMe3)2}(thf)2] ( 3 ) and [{N^N}Ba{CH(SiMe3)2}(thf)2] ( 10 ) are the first efficient Ba‐based precatalysts for intermolecular hydroamination and hydrophosphination, and display activity values that are above those reported so far. The potential of the precatalysts for C? N and C? P bond formation is detailed and a rare cyclohydroamination–intermolecular hydroamination “domino” sequence is presented.  相似文献   

20.
Heteroleptic silylamido complexes of the heavier alkaline earth elements calcium and strontium containing the highly fluorinated 3‐phenyl hydrotris(indazolyl)borate {F12‐Tp4Bo, 3Ph}? ligand have been synthesized by using salt metathesis reactions. The homoleptic precursors [Ae{N(SiMe3)2}2] (Ae=Ca, Sr) were treated with [Tl(F12‐Tp4Bo, 3Ph)] in pentane to form the corresponding heteroleptic complexes [(F12‐Tp4Bo, 3Ph)Ae{N(SiMe3)2}] (Ae=Ca ( 1 ); Sr ( 3 )). Compounds 1 and 3 are inert towards intermolecular redistribution. The molecular structures of 1 and 3 have been determined by using X‐ray diffraction. Compound 3 exhibits a Sr ??? MeSi agostic distortion. The synthesis of the homoleptic THF‐free compound [Ca{N(SiMe2H)2}2] ( 4 ) by transamination reaction between [Ca{N(SiMe3)2}2] and HN(SiMe2H)2 is also reported. This precursor constitutes a convenient starting material for the subsequent preparation of the THF‐free complex [(F12‐Tp4Bo, 3Ph)Ca{N(SiMe2H)2}] ( 5 ). Compound 5 is stabilized in the solid state by a Ca???β‐Si?H agostic interaction. Complexes 1 and 3 have been used as precatalysts for the intramolecular hydroamination of 2,2‐dimethylpent‐4‐en‐1‐amine. Compound 1 is highly active, converting completely 200 equivalents of aminoalkene in 16 min with 0.50 mol % catalyst loading at 25 °C.  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