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 46 毫秒
1.
Synthesis and Molekular Structures of N‐substituted Diethylgallium‐2‐pyridylmethylamides (2‐Pyridylmethyl)(tert‐butyldimethylsilyl)amine ( 1a ) and (2‐pyridylmethyl)‐di(tert‐butyl)silylamine ( 1b ) form with triethylgallane the corresponding red adducts 2a and 2b via an additional nitrogen‐gallium bond. These oily compounds decompose during distillation. Heating under reflux in toluene leads to the elimination of ethane and the formation of the red oils of [(2‐pyridylmethyl)(tert‐butyldimethylsilyl)amido]diethylgallane ( 3a ) and [(2‐pyridylmethyl)‐di(tert‐butyl)silylamido]diethylgallane ( 3b ). In order to investigate the thermal stability solvent‐free 3a is heated up to 400 °C. The elimination of ethane is observed again and the C‐C coupling product N, N′‐Bis(diethylgallyl)‐1, 2‐dipyridyl‐1, 2‐bis(tert‐butyldimethylsilyl)amido]ethan ( 4 ) is found in the residue. Substitution of the silyl substituents by another 2‐pyridylmethyl group and the reaction of this bis(2‐pyridylmethyl)amine with GaEt3 yield triethylgallane‐diethylgallium‐bis(2‐pyridylmethyl)amide ( 5 ). The metalation product adds immediately another equivalent of triethylgallane regardless of the stoichiometry. The reaction of GaEt3 with 2‐pyridylmethanol gives quantitatively colorless 2‐pyridylmethanolato diethylgallane ( 6 ).  相似文献   

2.
Metallation of N‐(diphenylphosphanyl)(2‐pyridylmethyl)amine with n‐butyllithium in toluene yields lithium N‐(diphenylphosphanyl)(2‐pyridylmethyl)amide ( 1 ), which crystallizes as a tetramer. Transamination of N‐(diphenylphosphanyl)(2‐pyridylmethyl)amine with an equimolar amount of Sn[N(SiMe3)2]2 leads to the formation of monomeric bis(trimethylsilyl)amido tin(II) N‐(diphenylphosphanyl)(2‐pyridylmethyl)amide ( 2 ). The addition of another equivalent of N‐(diphenylphosphanyl)(2‐pyridylmethyl)amine gives homoleptic tin(II) bis[N‐(diphenylphosphanyl)(2‐pyridylmethyl)amide] ( 3 ). In these complexes the N‐(diphenylphosphanyl)(2‐pyridylmethyl)amido groups act as bidentate bases through the nitrogen bases. At elevated temperatures HN(SiMe3)2 is liberated from bis(trimethylsilyl)amido tin(II) N‐(diphenylphosphanyl)(2‐pyridylmethyl)amide ( 2 ) yielding mononuclear tin(II) 1,2‐dipyridyl‐1,2‐bis(diphenylphosphanylamido)ethane ( 4 ) through a C–C coupling reaction. The three‐coordinate tin(II) atoms of 2 and 4 adopt trigonal pyramidal coordination spheres.  相似文献   

3.
(2‐Pyridylmethyl)(tert‐butyldimethylsilyl)amine ( 1 ) can be lithiated once or twice yielding lithium (2‐pyridylmethyl)(tert‐butyldimethylsilyl)amide ( 2 ) and dilithium (2‐pyridylmethanidyl)(tert‐butyldimethylsilyl)amide ( 3 ), respectively. The oxidation of 3 with white phosphorus yields dilithium 1,2‐dipyridyl‐1,2‐bis(tert‐butyldimethylsilylamido)ethane ( 4 ) which crystallizes after partial hydrolysis as an adduct of the form 2 · 4 .  相似文献   

4.
The metallation of (2‐pyridylethyl)(tert‐butyldimethylsilyl)amine ( 1 ) with dimethylzinc yields quantitatively dimeric methylzinc (2‐pyridylethyl)(tert‐butyldimethylsilyl)amide ( 2 ). Hydrolysis reactions lead to the precipitation of trakis(methylzinc) bis{2‐pyridylethyl)(tert‐butyldimethylsilyl}amide](μ4‐oxide)] ( 3 ).  相似文献   

5.
Synthesis of a Hexanuclear Calcium–Phosphorus‐Cage The metalation of tri(tert‐butyl)silylphosphane with calcium bis[bis(trimethylsilyl)amide] yields the dimer {(Me3Si)2N–Ca(THF)[μ‐P(H)SitBu3]}2 ( 1 ). In THF monomerization occurs and dismutation reactions lead to the homoleptic compounds, namely (THF)2Ca[N(SiMe3)2]2 and (THF)4Ca[P(H)SitBu3]2. In toluene, 1 undergoes dismutation reactions, bis(tetrahydrofuran)calcium bis[bis(trimethylsilyl)amide] is regained and [(Me3Si)2N–Ca(THF)]2Ca[P(H)SitBu3]4 ( 2 ) precipitates. At raised temperatures, 2 undergoes a homometallic metalation with the loss of two equivalents of HN(SiMe3)2 and dimerizes. The thus formed cage compound (THF)2Ca6[PSitBu3]4[P(H)SitBu3]4 ( 3 ) with a central Ca4P4 heterocubane moiety crystallizes upon cooling of the toluene solution. The molecular structures of 2 and 3 were determined.  相似文献   

6.
Crystal Structures and Spectroscopic Properties of 2λ3‐Phospha‐1, 3‐dionates and 1, 3‐Dionates of Calcium ‐ Comparative Studies on the 1, 3‐Diphenyl and 1, 3‐Di(tert‐butyl) Derivatives A hydrogen‐metal exchange between dibenzoylphosphane and calcium carbide in tetrahydrofuran (THF) followed by addition of the ligand 1, 3, 5‐trimethyl‐1, 3, 5‐triazinane (TMTA) furnishes the binuclear complex bis[(tmta‐N, N′, N″)calcium bis(dibenzoylphosphanide)] ( 1a ) co‐crystallizing with benzene. Similarly, reaction of bis(2, 2‐dimethylpropionyl)phosphane with bis(thf‐O)calcium bis[bis(trimethylsilyl)amide] in 1, 2‐dimethoxyethane (DME) gives bis(dme‐O, O′)calcium bis[bis(2, 2‐dimethylpropionyl)phosphanide] ( 1b ) in high yield. The carbon analogues 1, 3‐diphenylpropane‐1, 3‐dione (dibenzoylmethane) or 2, 2, 6, 6‐tetramethylheptane‐3, 5‐dione (dipivaloylmethane) and bis(thf‐O)calcium bis[tris(trimethylsilylmethyl)zincate] in DME afford bis(dme‐O, O′)calcium bis(dibenzoylmethanide) ( 2a ) and the binuclear complex (μ‐dme‐O, O′)bis[(dme‐O, O′)calcium bis(dipivaloylmethanide)] ( 2b ), respectively. Dialkylzinc formed during the metalation reaction shows no reactivity towards the 1, 3‐dionates 2a and 2b . Finally, from the reaction of the unsymmetrically substituted ligand 2‐(methoxycarbonyl)cyclopentanone and bis(thf‐O)calcium bis[bis(trimethylsilyl)amide] in toluene, the trinuclear complex 3 is obtained, co‐crystallizing with THF. The β‐ketoester anion bridges solely via the cyclopentanone unit.  相似文献   

7.
The metalation of HP(SiMe3)2 with Y[CH(SiMe3)2]3 gives the homoleptic {Y[P(SiMe3)2]3}2 (1) which crystallizes from toluene in the monoclinic space group P21/c. The yttrium atoms are in a distorted tetrahedral environment with Y‐P bond lengths of 267.7 and 284.8 pm to the terminal and bridging substituents, respectively. The metathesis reaction of [1, 3‐(Me3Si)2C5H3]2YCl with KPSitBu3 yields (tetrahydrofuran‐O)‐1, 1', 3, 3'‐tetrakis(trimethylsilyl)yttrocene‐tri(tert‐butyl)silylphosphanide ( 2 ). The molecular structure of 2 in solution was deduced by NMR spectroscopy and X‐ray crystallography. The coupling constants 1J(Y, P) and 1J(P, H) show values of 144.0 Hz and 201.0 Hz, respectively.  相似文献   

8.
Synthesis, Spectroscopic Characterization, and Molecular Structures of Selected Lewis‐Base Adducts of the Alkali Metal Tri(tert‐butyl)silylphosphanides The metalation of tri(tert‐butyl)silylphosphane with butyllithium and the bis(trimethylsilyl)amides of sodium, potassium, and rubidium yields quantitatively the corresponding alkali metal tri(tert‐butyl)silylphosphanides, which crystallize after addition of appropriate Lewis‐bases as dimeric (DME)LiP(H)SitBu3 ( 1 ), chain‐like (DME)NaP(H)SitBu3 ( 2 ), monomeric ([18]Krone‐6)KP(H)SitBu3 ( 3 ), and dimeric (TMEDA)1.5RbP(H)SitBu3 ( 4 ). The reaction of H2PSitBu3 with cesium bis(trimethylsilyl)amide at room temperature gives monocyclic and tetrameric cesium tri(tert‐butyl)silylphosphanide ( 5 ) with two additional coordinated CsN(SiMe3)2 molecules. At 80 °C this complex reacts with excess of phosphane to the tetrameric toluene adduct (η6‐Toluol)CsP(H)SitBu3 ( 6 ) which contains a central Cs4P4‐heterocubane fragment. The constitution of these compounds was verified by X‐ray structure determinations.  相似文献   

9.
Synthesis and Characterization of Hetero-bimetallic Bis(trimethylsilyl)phosphanides of Barium and Tin The reaction of barium bis[bis(trimethylsilyl)amide] with one equivalent of bis(trimethylsilyl)phosphane in 1,2-dimethoxyethane (dme) yields the heteroleptic dimeric (dme)barium bis(trimethylsilyl)amide bis(trimethylsilyl)phosphanide. This colorless compound crystallizes in the monoclinic space group P21/n with a = 1 259.1(3), b = 1 822.7(4), c = 1 516.1(3) pm, β = 110.54(3)° and Z = 4. The central moiety of the centrosymmetric molecule is the planar Ba2P2-cycle with Ba? P-bond lengths of 329 and 334 pm. In the presence of bis[bis(trimethylsilyl)amino]stannylene hetero-bimetallic bis(trimethylsilyl)phosphanides of tin(II) and barium are isolated. If the reaction of Ba[N(SiMe3)2]2 and Sn[N(SiMe3)2]2 in the molar ratio of 1:2 with six equivalents of HP(SiMe3)2 is performed in toluene, barium bis{tin(II)-tris[bis(trimethylsilyl)phosphanide]} can be isolated. This compound crystallizes in the orthorhombic space group P212121 with a = 1 265.1(1), b = 2 290.1(3), c = 2 731.9(3) pm and Z = 4. The anions {Sn[P(SiMe3)2]3}? bind as two-dentate ligands to the barium atom which shows the extraordinary low coordination number of four. The addition of tetrahydrofuran (thf) to the above mentioned reaction solution leads to the elimination of tris(trimethylsilyl)phosphane and the formation of thf complexes of barium bis{tin(II)-bis(trimethylsilyl)phosphanide-trimethylsilylphosphandiide}. The derivative crystallizes from toluene in the monoclinic space group P21/c with a = 1 301.9(2), b = 2 316.3(3), c = 3 968.7(5) pm, β = 99.29(1)° and Z = 8.  相似文献   

10.
Synthesis of Trimethylsilyl Substituted Polyhedra of Calcium, Tin(II), and Phosphorus The reaction of calcium-bis[bis(trimethylsilyl)amide] with bis(trimethylsilyl)phosphane in thf yields the heteroleptic, dimeric (tetrahydrofuran-O)calcium-bis(trimethylsilyl)amidebis(trimethylsilyl)phosphanide 1 (triclinic, P 1 , a = 1066,6(2), b = 1141,3(2), c = 1226,6(2)pm, α = 97,78(3)°, β = 107,47(3)°, γ = 101,12(3)°, Z = 1 dimer). The bridging phosphanide-substituent displays with Ca? P bond lengths of 292,6 and 300,5 pm a distortion of the four-membered Ca2P2-cycle. The reaction with another equivalent of HP(SiMe3)2 in thf leads to the formation of tetrakis(tetrahydrofuran-O)calcium-bis[bis(trimethylsilyl)phosphanide] 2 mit Ca? P distances of 292 pm (monoclinic, P21/c, a = 1626,0(3), b = 1295,3(4), c = 2039,5(5) pm, β = 102,60(2)°, Z = 4). The performance of the reaction in the presence of bis[bis(trimethylsilyl)amino]stannylene yields heterobimetallic compounds with a central polyhedron of Ca-, Sn- and P-atoms. Dependent on the Sn/Ca ratio the isolation of tris(trimethylsilyl)phosphane as well as bis[tris(tetrahydrofuran-O)calcium]-ditin(II)-tetrakis(μ3-trimethylsilylphosphandiide) 3 with a central dicalcia-distanna-tetraphosphacubane-fragment or (thf)2CaSn2[μ-P(SiMe3)2]23-PSiMe3]2 4 (orthorhombic, Pnma, a = 2247,7(2), b = 1868,9(1), c = 1168,0(1) pm, Z = 4), respectively, succeeds. The Ca? P distances lie at 291 pm.  相似文献   

11.
Two series of new dinuclear rare‐earth metal alkyl complexes supported by indolyl ligands in novel μ‐η211 hapticities are synthesized and characterized. Treatment of [RE(CH2SiMe3)3(thf)2] with 1 equivalent of 3‐(tBuN?CH)C8H5NH ( L1 ) in THF gives the dinuclear rare‐earth metal alkyl complexes trans‐[(μη211‐3‐{tBuNCH(CH2SiMe3)}Ind)RE(thf)(CH2SiMe3)]2 (Ind=indolyl, RE=Y, Dy, or Yb) in good yields. In the process, the indole unit of L1 is deprotonated by the metal alkyl species and the imino C?N group is transferred to the amido group by alkyl CH2SiMe3 insertion, affording a new dianionic ligand that bridges two metal alkyl units in μη211 bonding modes, forming the dinuclear rare‐earth metal alkyl complexes. When L1 is reduced to 3‐(tBuNHCH2)C8H5NH ( L2 ), the reaction of [Yb(CH2SiMe3)3(thf)2] with 1 equivalent of L2 in THF, interestingly, generated the trans‐[(μη211‐3‐{tBuNCH2}Ind)Yb(thf)(CH2SiMe3)]2 (major) and cis‐[(μη211‐3‐{tBuNCH2}Ind)Yb(thf)(CH2SiMe3)]2 (minor) complexes. The catalytic activities of these dinuclear rare‐earth metal alkyl complexes for isoprene polymerization were investigated; the yttrium and dysprosium complexes exhibited high catalytic activities and high regio‐ and stereoselectivities for isoprene 1,4‐cis‐polymerization.  相似文献   

12.
The reaction of dimethylzinc and tri(tert‐butyl)silylphosphane in toluene yielded dimeric methylzinc tri(tert‐butyl)silylphosphanide ( 1 ) which crystallized tetrameric. Compound 1 was deprotonated with sodium in DME and the solvent‐separated dimeric ion pair [(dme)3Na]+ [(dme)Na(MeZn)2(μ‐PSitBu3)2]? ( 2 ) was isolated. The reaction of 1 in THF with two equivalents of potassium and one equivalent of tri(tert‐butyl)silylphosphane gave dimeric [{tBu3Si(H)P}{(thf)2K}2(MeZn)(PSitBu3)]2 ( 3 ). Both of these phosphanylzincates contain Zn2P2 cycles with Zn‐P bond lengths of approximately 237 pm, whereas in 1 larger Zn‐P bond lengths of 248.5 pm were found due to the larger coordination numbers of the phosphorus and zinc atoms.  相似文献   

13.
We report that the formation of μ‐oxo diferric compounds from O2 and FeCl2 complexes within the tris(2‐pyridylmethyl)amine series (N. K. Thallaj et al. Chem. Eur. J., 2008 , 14, 6742–6753) involves coordination of O2 to the metal centre and that this reaction occurs following initial dissociation of the bound equatorial chloride anion. We also report evidence of the formation of a reduced form of dioxygen by an inner‐sphere mechanism, thus leading to modification of the ligand. The solid‐state structures of [FeCl2L] complexes (L1=mono(α‐pivalamidopyridylmethyl)bis(2‐pyridylmethyl)amine, L2=mono(α‐pivalesteropyridylmethyl)bis(2‐pyridylmethyl)amine, L3=bis(α‐pivalamidopyridylmethyl)mono(2‐pyridylmethyl)amine are described, and spectroscopic data support the structural retention in solution. In [FeCl2L3], the two amide hydrogen atoms stabilise the equatorial chloride anion in such a way that its exchange by a weak ligand is impossible: [FeCl2L3] is perfectly oxygen‐stable. In [FeCl2L2], the equatorial chloride anion is completely free to move and coordination of O2 can take place. The reaction product with [FeCl2L2] is a μ‐oxo diferric complex in which the ester function has been transformed into a phenol group. This conversion can be seen as a hydrolysis reaction in basic medium, hence supporting the initial formation of a reduced form of dioxygen in the medium. Complex [FeCl2L1] exhibits a very weak reactivity with O2, in line with a semistabilised equatorial chloride counteranion.  相似文献   

14.
The behavior of the first aminophenolate catalysts of the large alkaline earth metals (Ae) [(LOi)AeN(SiMe2R)2(thf)x] (i=1–4; Ae=Ca, Sr, Ba; R=H, Me; x=0–2) for the cyclohydroamination of terminal aminoalkenes is discussed. The complexes [(BDI)AeN(SiMe2H)2(thf)x] (Ae=Ca, Sr, Ba, x=1–2; (BDI)H=H2C[C(Me)N‐2,6‐(iPr)2C6H3]2)) and [(BDI)CaN(SiMe3)2(thf)] supported by the β‐diketiminate (BDI)? ligand have also been employed for comparative and mechanistic considerations. The catalytic performances decrease in the order Ca>Sr?Ba, which is the opposite trend to that previously observed during the intermolecular hydroamination of activated alkenes catalyzed by the same alkaline‐earth metal complexes. Catalyst efficacy increases when the chelating and donating ability of the aminophenolate ligands decreases. For given metals and ancillary scaffolds, disilazide catalysts that incorporate the N(SiMe3)2? amido group outclass their congeners containing the N(SiMe2H)2? amide owing to the lower basicity of the N(SiMe2H)2? with respect to the N(SiMe3)2? group, and also because Ae–N(SiMe2H)2 catalysts suffer from irreversible deactivation through the dehydrogenative coupling of amine and hydrosilane moieties. This deactivation process takes place at 25 °C in the case of [(LOi)AeN(SiMe2H)2(thf)x] phenolate complexes and occurs even with the related [(BDI)AeN(SiMe2H)2(thf)x] complex, albeit under conditions harsher than those required for effective cyclohydroamination catalysis. A mechanistic scenario for cyclohydroamination catalyzed by [(LX)AeN(SiMe2H)2(thf)x] complexes ((LX)?=(LOi)? or (BDI)?) is proposed. Although beneficial for the synthesis of Ae heteroleptic complexes able to resist deleterious Schlenk‐type equilibria, the use of the N(SiMe2H)2? is prejudicial to catalytic activity in the case of catalyzed transformations that involve reactive amine (and potentially other) substrates. Mechanistic and kinetic investigations further illustrate the interplay between the catalytic activity, operative mechanism, and identity of the metal, ancillary ligand, and amido group. These studies suggest that the widely accepted mechanism for cyclohydroamination reactions cannot be extended systematically to all alkaline‐earth catalysts. The [(BDI)CaN(SiMe2H)2{H2NCH2C(CH3)2CH2CH?CH2}2] complex, the first Ca–aminoalkene adduct structurally characterized, was prepared quantitatively and essentially behaves like [(BDI)CaN(SiMe2H)(thf)], thus serving as a model compound for mechanistic studies, as illustrated during stoichiometric reactions monitored by 1H NMR spectroscopy.  相似文献   

15.
Transmetallation of Tin(II) in [Sn(μ3‐PSitBu3)]4 by Barium – from Sn4P4 Heterocubane Structures to Heterobinuclear Cage Compounds with a Central BanSn4?nP4 Heterocubane Polyhedron (n = 1, 2 and 3) For the preparation of compounds of the type [BanSn4?n(PSitBu3)4] (n = 1 ( 2 ), 2 ( 3 ) and 3 ( 4 )) two synthetic routes are applicable: in the transmetallation reaction homometallic [Sn4(PSitBu3)4] ( 1 ) reacts with barium metal and in a deprotonation reaction (metallation) tri(tert‐butyl)silylphosphane reacts simultaneously with (thf)2Ba[N(SiMe3)2]2 and Sn[N(SiMe3)2]2. During the transmetallation reaction mixtures of the heterobimetallic cage compounds 2 to 4 are obtained, however, analytically pure compounds 2 and 3 are accessible by the metallation reaction. Compound 4 is formed as a minor product together with 3 . Due to the larger Ba‐P bond lengths compared to the Sn‐P values the substitution of tin by barium leads to strong distortions of the heterocubane moiety. With NMR‐spectroscopic experiments one could show that all the above mentioned compounds form BanSn4?nP4 heterocubane cage structures.  相似文献   

16.
The synthesis of two novel titanium carbene complexes from the bis(thiophosphinoyl)methanediide geminal dianion 1 (SCS2?) is described. Dianion 1 reacts cleanly with 0.5 equivalents of [TiCl4(thf)2] to afford the bis‐carbene complex [(SCS)2Ti] ( 2 ) in 86 % yield. The mono‐carbene complex [(SCS)TiCl2(thf)] ( 3 ) can also be obtained by using an excess of [TiCl4(thf)2]. The structures of 2 and 3 are confirmed by X‐ray crystallography. A strong nucleophilic reactivity towards various electrophiles (ketones and aldehydes) is observed. The reaction of 3 with N,N′‐dicyclohexylcarbodiimide (DCC) and phenyl isocyanate leads to the formation of two novel diphosphinoketenimines 8 a and 8 b . The bis‐titanium guanidinate complex 9 is trapped as the by‐product of the reaction with DCC. The X‐ray crystal structures of 8 a and 9 are presented. The mechanism of the reaction between complex 3 and DCC is rationalized by DFT studies.  相似文献   

17.
The reaction of potassium 2,5‐bis[N‐(2,6‐diisopropylphenyl)iminomethyl]pyrrolyl [(dip2‐pyr)K] with the borohydrides of the larger rare‐earth metals, [Ln(BH4)3(thf)3] (Ln=La, Nd), afforded the expected products [Ln(BH4)2(dip2‐pyr)(thf)2]. As usual, the trisborohydrides reacted like pseudohalide compounds forming KBH4 as a by‐product. To compare the reactivity with the analogous halides, the dimeric neodymium complex [NdCl2(dip2‐pyr)(thf)]2 was prepared by reaction of [(dip2‐pyr)K] with anhydrous NdCl3. Reaction of [(dip2‐pyr)K] with the borohydrides of the smaller rare‐earth metals, [Sc(BH4)3(thf)2] and [Lu(BH4)3(thf)3], resulted in a redox reaction of the BH4? group with one of the Schiff base functions of the ligand. In the resulting products, [Ln(BH4){(dip)(dip‐BH3)‐pyr}(thf)2] (Ln=Sc, Lu), a dinegatively charged ligand with a new amido function, a Schiff base, and the pyrrolyl function is bound to the metal atom. The by‐product of the reaction of the BH4? anion with the Schiff base function (a BH3 molecule) is trapped in a unique reaction mode in the coordination sphere of the metal complex. The BH3 molecule coordinates in an η2 fashion to the metal atom. The rare‐earth‐metal atoms are surrounded by the η2‐coordinated BH3 molecule, the η3‐coordinated BH4? anion, two THF molecules, and the nitrogen atoms from the Schiff base and the pyrrolyl function. All new compounds were characterized by single‐crystal X‐ray diffraction. Low‐temperature X‐ray diffraction data at 6 K were collected to locate the hydrogen atoms of [Lu(BH4){(dip)(dip‐BH3)‐pyr}(thf)2]. The (DIP2‐pyr)? borohydride and chloride complexes of neodymium, [Nd(BH4)2(dip2‐pyr)(thf)2] and [NdCl2(dip2‐pyr)(thf)]2, were also used as Ziegler–Natta catalysts for the polymerization of 1,3‐butadiene to yield poly(cis‐1,4‐butadiene). Very high activities and good cis selectivities were observed by using each of these complexes as a catalyst in the presence of various cocatalyst mixtures.  相似文献   

18.
Reaction of the tripodal trilithium triamide [MeSi{SiMe2N(Li)(p‐Tol)}3(thf)2] ( 1 ) with SbCl3 in toluene gave the corresponding triaminostibine [MeSi{SiMe2N(p‐Tol)}3Sb] ( 2 ) in good yield. Its [2,2,2]bicyclooctane‐related cage structure, comprising the trisilylsilane unit and the triamidostibine fragment, was established by a single crystal X‐ray structure analysis: Space group P1, Z = 2, lattice dimensions at 293 K: a = 8.645(4), b = 10.029(5), c = 19.678(9) Å, α = 98.50(3)°, β = 97.46(3)°, γ = 94.80(3)°, R = 0.0216.  相似文献   

19.
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   

20.
The treatment of the recently reported potassium salt (S)‐N,N′‐bis‐(1‐phenylethyl)benzamidinate ((S)‐KPEBA) and its racemic isomer (rac‐KPEBA) with anhydrous lanthanide trichlorides (Ln=Sm, Er, Yb, Lu) afforded mostly chiral complexes. The tris(amidinate) complex [{(S)‐PEBA}3Sm], bis(amidinate) complexes [{Ln(PEBA)2(μ‐Cl)}2] (Ln=Sm, Er, Yb, Lu), and mono(amidinate) compounds [Ln(PEBA)(Cl)2(thf)n] (Ln=Sm, Yb, Lu) were isolated and structurally characterized. As a result of steric effects, the homoleptic 3:1 complexes of the smaller lanthanide atoms Yb and Lu were not accessible. Furthermore, chiral bis(amidinate)–amido complexes [{(S)‐PEBA}2Ln{N(SiMe3)2}] (Ln=Y, Lu) were synthesized by an amine‐elimination reaction and salt metathesis. All of these chiral bis‐ and tris(amidinate) complexes had additional axial chirality and they all crystallized as diastereomerically pure compounds. By using rac‐PEBA as a ligand, an achiral meso arrangement of the ligands was observed. The catalytic activities and enantioselectivities of [{(S)‐PEBA}2Ln{N(SiMe3)2}] (Ln=Y, Lu) were investigated in hydroamination/cyclization reactions. A clear dependence of the rate of reaction and enantioselectivity on the ionic radius was observed, which showed higher reaction rates but poorer enantioselectivities for the yttrium compound.  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