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, Properties, and Reaction Behaviour of 2-(Dimethylaminomethyl)phenyl- and 8-(Dimethylamino)naphthylsubstituted Lithium Hydridosilylamides – Formation of Silanimines by Elimination of Lithium Hydride The hydridosilylamines Ar(R)Si(H)–NHR′ ( 2 a : Ar = 2-Me2NCH2C6H4, R = Me, R′ = CMe3; 2 b : Ar = 2-Me2NCH2C6H4, R = Ph, R′ = CMe3; 2 c : Ar = 2-Me2NCH2C6H4, R = Me, R′ = SiMe3; 2 d : Ar = 8-Me2NC10H6, R = Me, R′ = CMe3; 2 e : Ar = 8-Me2NC10H6, R = Ph, R′ = CMe3; 2 f : Ar = 8-Me2NC10H6, R = Me, R′ = SiMe3) have been synthesized from the appropriate chlorosilanes Ar(R)SiHCl either by reaction with the stoichiometric amount of Me3CNHLi ( 2 a , 2 b , 2 d , 2 e ) or by coammonolysis in liquid NH3 with chlorotrimethylsilane in molar ratio 1 : 3 ( 2 c , 2 f ). Treatment of 2 a–2 f with n-butyllithium in equimolar ratio in n-hexane resulted in the lithiumhydridosilylamides Ar(R)Si(H)–N(Li)R′ 3 a–3 f . The frequencies of the Si–H stretching vibration and 29Si–1H coupling constants in the amides are smaller than in the analogous amines indicating a higher hydride character for the hydrogen atom of the Si–H group in the amides compared to the amines. Results of NMR spectroscopic studies point to the existence of a (Me2)N → Si coordination bond in the 8-(dimethylamino)naphthyl-substituted amines and amides. The amides 3 a–3 c are stable under refluxing in m-xylene. At the same conditions 3 d and 3 e eliminate LiH and the silanimines 8-Me2NC10H6(R)Si=NCMe3 ( 4 d : R = Me, 4 e : R = Ph) are formed. The amides 3 a–3 d und 3 f react with chlorotrimethylsilane in THF to give the corresponding N-substitution products Ar(R)Si(H)–N(SiMe3)R′ 6 a–6 d and 6 f in good yields. 4 d is formed as a byproduct in the reaction of 3 d with chlorotrimethylsilane. In n-hexane and m-xylene these amides are little reactive opposite to chlorotrimethylsilane. 6 a–6 d and 6 f are obtained in very small amounts. In the case of 3 d besides the N-substitution product 6 d the silanimine 4 d is obtained. In contrast to chlorotrimethylsilane the amides 3 a and 3 f react well with chlorodimethylsilane in m-xylene producing 2-Me2NCH2C6H4(H) SiMe–N(SiHMe2)CMe3 ( 7 a ) and 8-Me2NC10H6(H)SiMe–N(SiHMe2)SiMe3 ( 7 f ).  相似文献   

2.
The reaction of the stannylene R2Sn : (R = 2-tBu-4,5,6-Me3C6H) with R′2Sn (R′ = Si(SiMe3)3) proceeds with substituent exchange to afford the heteroleptic stannylene RR′Sn : which, in the solid state, forms the distannene RR′Sn = SnRR′ ( 7 ). The X-ray structure analysis of 7 reveals a trans-bent arrangement of the substituents with a large fold angle of 44.9° and an Sn–Sn double bond length of 279.14(4) pm.  相似文献   

3.
李悦生 《高分子科学》2011,29(5):627-633
Mono salicylaldiminato vanadium(Ⅲ) complexes(1a-1f)[RN = CH(ArO)]VCl2(THF)2(Ar = C6H4(1a-1e),R = Ph,1a;R = p-CF3Ph,1b;R = 2,6-Me2Ph,1c;R = 2,6-iPr2Ph,1d;R = cyclohexyl,1e;Ar = C6H2tBu2(2,4),R = 2,6-iPr2Ph, 1f) and bis(salicylaldiminato) vanadium(Ⅲ) complexes(2a-2f)[RN = CH(ArO)]2VCl(THF)x(Ar = C6H4(2a-2e),x = 1 (2a-2e),R = Ph,2a;R =p-CF3Ph,2b;R = 2,6-Me2Ph,2c;R = 2,6-iPr2Ph,2d;R = cyclohexyl,2e;Ar = C6H2tBu2(2,4),R = 2,6-iPr2Ph,x = 0,2f) have been evaluated as the active catalysts for ethylene/1-hexene copolymerization in the presence of Et2AlCl.The ligand substitution pattern and the catalyst structure model significantly influenced the polymerization behaviors such as the catalytic activity,the molecular weight and molecular weight distribution of the copolymers etc.The highest catalytic activity of 8.82 kg PE/(mmolV·h) was observed for vanadium catalyst 2d with two 2,6-diisopropylphenyl substituted salicylaldiminato ligands.The copolymer with the highest molecular weight was obtained by using mono salicylaldiminato vanadium catalyst 1f having ligands with tert-butyl at the ortho and para of the aryloxy moiety.  相似文献   

4.
Homoleptic Amides of Zinc, Cadmium, and Mercury ZnCl2, CdCl2 and HgCl2 react with the lithium salts ( 1 a–5 a ) of the sterically demanding secundary amines HN(SiMe3)Ph ( 1 ), HN(SiMe3)C6H3Me2‐2,6 ( 2 ), HN(SiMe3)C6H3iPr2‐2,6 ( 3 ), HN(SiMe3)C6H3tBu2‐2,5 ( 4 ), and HN(SiMe2NMe2)C6H3iPr2‐2,6 ( 5 ) yielding the corresponding homoleptic metal amides Zn[N(SiMe2R′)R]2 ( 1 b–5 b ), Cd[N(SiMe2R′)R]2 ( 1 c , 5 c ), and Hg[N(SiMe2R′)R]2 ( 1 d–5 d ), respectively. Except the dimeric {Zn[N(SiMe3)Ph]2}2 ( 1 b ), all complexes are monomeric. The compounds were characterized by elemental analyses, molecular weight determinations, NMR and mass spectra. Furthermore, the zinc amides ( 1 b–5 b ) and the mercury amides 1 d–3 d and 5 d were characterized by single crystal X‐ray structure analysis. Except 1 b and 5 b , they show a linear N–M–N arrangement.  相似文献   

5.
The bis(amidodimethyl)disiloxane antimony chlorides Sb(NONR)Cl (NONR=[O(SiMe2NR)2]2−; R=tBu, Ph, 2,6-Me2C6H3=Dmp, 2,6-iPr2C6H3=Dipp, 2,6-(CHPh2)2-4-tBuC6H2=tBu-Bhp) are reduced to SbII and SbI species by using MgI reagents, [Mg(BDIR′)]2 (BDI=[HC{C(Me)NR′}2]; R′=2,4,6-Me3C6H2=Mes, Dipp). Stoichiometric reactions with Sb(NONR)Cl (R=tBu, Ph) form dimeric SbII stibanes [Sb(NONR)]2, shown crystallographically to contain Sb−Sb single bonds. The analogous distibane with R=Dmp substituents has an exceptionally long Sb−Sb interaction and exhibits spectroscopic and reactivity properties consistent with radical character in solution. When R=Dipp, reductions with MgI reagents directly give distibenes [Sb(μ-NONDipp)Mg(BDIR′)(THF)n]2 (R′=Mes, n=1; R′=Dipp, n=0). Crystallographic analysis shows a trans-substitution of the Sb=Sb double bond, with bridging NONDipp-ligands between the SbI and MgII centres. An attempt to access the NONPh-analogue using the same protocol afforded the polystibide cluster Sb8[μ4,η2:2:2:2-Mg(BDIMes)]4, which co-crystallized with the ligand transfer product, [Mg(BDIMes)]2(μ-NONPh).  相似文献   

6.
The ring-opening Si-fluorination of a variety of azasilole derivatives cyclo-1-(iPr2Si)−4-X−C6H3−2-CH2NR ( 4 : R=2,6-iPr2C6H3, X=H; 4 a : R=2,4,6-Me3C6H2, X=H; 9 : R=2,6-iPr2C6H3, X=tBuMe2SiO; 10 : R=2,6-iPr2C6H3, X=OH; 13 : R=2,6-iPr2C6H3, X=HCCCH2O; 22 : R=2,6-iPr2C6H3, X=tBuMe2SiCH2O) with different 19F-fluoride sources was studied, optimized and the experience gained was used in a translational approach to create a straightforward 18F-labelling protocol for the azasilole derivatives [18F] 6 and [18F] 14 . The latter constitutes a potential clickable CycloSiFA prosthetic group which might be used in PET tracer development using Cu-catalysed triazole formation. Based on our findings, CycloSiFA has the potential to become a new entry into non-canonical labelling methodologies for radioactive PET tracer development.  相似文献   

7.
O-Halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines – Synthesis, Crystal Structure, and Reactions The substitution of halogenosilanes on lithiated N,O-bis(trimethylsilyl)-hydroxylamine in the molar ratio of 1 : 1 occurs on the oxygen atom. The O-halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines were prepared: RSiF2ON · (SiMe3)2 (R = CMe3 1 , CHMe2 2 , CH2C6H5 3 , C6H2(CMe3)3 4 ), RR′SiFON(SiMe3)2 (R = CMe3, R′ = C6H5 5 ; R = Me, R′ = C6H5 6 ; R = C6H2Me3, R′ = C6H2Me3 7 ; R = CH2C6H5, R′ = CH2C6H5 8 ; R = CHMe2, R′ = CHMe2 9 ; R = CMe3, R′ = CMe3 10 ), RSiCl2ON(SiMe3)2 (R = CMe3 11 ; R = Cl 12 ). The reaction of fluorosilanes with lithiated N,O-bis(trimethylsilyl)hydroxylamine in the molar ratio of 1 : 2 leads to the formation of O,O′-fluorosilyl-bis[N,N-bis(trimethylsilyl)hydroxylamines]: RSiF[ON(SiMe3)2]2 (R = CMe3 13 ; R = C6H5 14 ). 13 could be prepared in the reaction of 1 with LiON(SiMe3)2. Lithiated dimethylketonoxime reacts with 1 to Me2C=NOSiRF–ON(SiMe3)2 [R = CMe3 ( 15 )]. The first crystal structure of a tris(silyl)hydroxylamine ( 4 ) is shown. The angle at the nitrogen prove a pyramidal geometry.  相似文献   

8.
The behavior of different anilines H2NC6H4R (R = o-Me, p-Me, o-, m- and p? i Pr, p-OMe, p-CO2Et) and 2,6-Me2C6H3NH2 towards trihalophosphoranes was studied. 2,6-Me2C6H3NH2 failed to form the diaminophosphonium salt [Ph2PNH(2,6-Me2C6H3)2]Br, and the aminophosphine oxide Ph2(2,6-Me2C6H3NH)PO was the only isolated product. Both o- and p-toluidine gave the corresponding diaminophosphonium salts; however in the case of o-toluidine, the yield was low and a mixture with the respective aminophosphine oxide was observed. Anilines containing methoxy and ethoxycarbonyl groups in para-position form the diaminophosphonium salts in reasonable yields.  相似文献   

9.
The diaminosilylene tBuNCH2CH2NtBuSi: reacted with the diaminogermylenes RNCH2CH2NRGe: R = 2,6-Me2C6H3, iPr, by silylene insertion into one of the Ge–N bonds to furnish the aminosilylgermylenes 8 , R = 2,6-Me2C6H3, and 9 , R = iPr. The X-ray structure analyses of these compounds revealed that 8 remains monomeric in the crystal with weak Ge … Ge interactions to the germanium atom of a neighbouring germylene molecule, whereas 9 dimerizes to give the strongly twisted (E)-1,2-diamino-1,2-disilyldigermene (E)- 10 with a long Ge–Ge double bond of 246 pm and a large trans-bent angle of 47.3°.  相似文献   

10.
The reactions of PhCH2SiMe3 ( 1 ), PhCH2SiMe2tBu ( 2 ), PhCH2SiMe2Ph ( 3 ), 3,5‐Me2C6H3CH2SiMe3 ( 4 ), and 3,5‐Me2C6H3CH2SiMe2tBu ( 5 ) with nBuLi in tetramethylethylenediamine (tmeda) afford the corresponding lithium complexes [Li(tmeda)][CHRSiMe2R′] (R, R′ = Ph, Me ( 6 ), Ph, tBu ( 7 ), Ph, Ph ( 8 ), 3,5‐Me2C6H3, Me ( 9 ), and 3,5‐Me2C6H3, tBu ( 10 )), respectively. The new compounds 5 , 7 , 8 , 9 and 10 have been characterized by 1H and 13C NMR spectroscopy, compounds 7 , 8 and 9 also by X‐ray structure analysis.  相似文献   

11.
The reaction of O,O′-diisopropylphosphoric acid isothiocyanate (iPrO)2P(O)NCS with 2-methylaniline 2-MeC6H4NH2, 2,6-dimethylaniline 2,6-Me2C6H3NH2, or 2,4,6-trimethylaniline 2,4,6-Me3C6H2NH2 leads to the N-phosphorylated thioureas RNHC(S)NHP(O)(OiPr)2 (R = 2-MeC6H4?, HLI ; 2,6-Me2C6H3?, HLII ; 2,4,6-Me3C6H2?, HLIII ). Reaction of the potassium salts of HLI III with Ni(II) in aqueous EtOH leads to [Ni(LI–III-N,S)2] ([NiLI–III 2 ]) chelate complexes. The compounds obtained were investigated by 1H, 31P{1H} NMR spectroscopy and microanalysis. The molecular structure of the thiourea HLIII was elucidated by single crystal X-ray diffraction analysis. Single crystal X-ray diffraction studies showed that HLIII forms both intra- and intermolecular hydrogen bonds, which in turn leads to the formation of polymeric chains. One of the intermolecular hydrogen bonds is of the type N?H…S. Moreover, the formation of intermolecular C?H…η6-phenyl interactions was established.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.  相似文献   

12.
Effect of para-substituents in the ethylene (E) copolymerization with 1-decene (DC), 1-dodecene (DD), and with 2-methyl-1-pentene (2M1P) using a series of Cp*TiCl2(O-2,6-iPr2-4-R-C6H2) [R=H ( 1 ), tBu ( 2 ), Ph ( 3 ), CHPh2 ( 4 ), CPh3 ( 5 ), SiMe3 ( 6 ), SiEt3 ( 7 ), and newly prepared 4-tBuC6H4 ( 8 ) and 3,5-Me2C6H3 ( 9 )]-MAO catalyst systems has been studied. The activities in these copolymerization reactions were affected by the para-substituent, and the SiMe3 ( 6 ), SiEt3 ( 7 ) and 3,5-Me2C6H3 ( 9 ) analogues showed the higher activities at 50 °C in the E copolymerization reactions with DC (1.06–1.44×106 kg-polymer/mol-Ti⋅h), DD (1.04–1.88×106 kg-polymer/mol-Ti⋅h) than the others, whereas no significant differences were observed in the comonomer incorporations. Complexes 6 and 7 also showed the higher activities at 50 °C in the E/2M1P copolymerization, and the 2M1P incorporation was affected by the para-substituent and the polymerization temperature; complex 9 showed better 2M1P incorporation at 25 °C.  相似文献   

13.
The thermal LiHal elimination of
- and
functional compounds provides a simple synthetic route to four-membered SiC and SiN rings. In attempts to inhibit dimerisation sterically, bulky silylmethyl and silylamino substituents were introduced (I–III). (Me3Si)3CSiF2R reacts with LiNHR′, 1,3- migration of a silyl group from carbon to the nitrogen (I, R′= 2,4,6-Me3C6H2) taking place. Substitution occurs for R′ = SiMe2CMe2, (II, III) only.Dichloro-bis(trimethylsilyl)methane reacts with halogenosilanes and lithium in THF to give bis(trimethylsilyl)-halogenosilaethanes (Me3Si)2CHSi(Hal)RR′; R= Me, R′ = N(SiMe3)2, IV, Hal = F; V, Hal = Cl. However a reductive THF cleavage accompanied by a silyl group migration to the oxygen occurs and 1-halogenosilyl-1- trimethylsilyl-5-trimethylsiloxi-pent-1-ene,(Me3Si)(RR′SiHal)CCH(CH2)3OSiMe3, Are The main products (VII–X) of these reactions. Disubstitution occurs with F3Si-i-Pr (VI). (Me3Si)3CSiFNHSiMe2CMe3 (II) reacts with C4H9Li in a molar ratio 12 to give an 1-aza-2,3-disilacyclobutane (XI), involving substitution, LiF elimination, and nucleophilic migration of a methanide ion of the unsaturated precusor.(Me3Si)2CHSiFMeN (2,4,6-Me3C6H2)SiMe3 cyclizes under comparable conditions in the reaction with MeLi via a methylene group of the mesityl group (XII).  相似文献   

14.
Reactions of Lithium Hydridosilylamides RR′(H)Si–N(Li)R″ with Chlorotrimethylsilane in Tetrahydrofuran and Nonpolar Solvents: N‐Silylation and/or Formation of Cyclodisilazanes The lithiumhydridosilylamides RR′(H)Si–N(Li)R″ ( 2 a : R = R′ = CHMe2, R″ = SiMe3; 2 b : R = R′ = Ph, R″ = SiMe3; 2 c : R = R′ = CMe3, R″ = SiMe3; 2 d : R = R′ = R″ = CMe3; 2 e : R = Me, R′ = Si(SiMe3)3, R″ = CMe3; 2 f – 2 h : R = R′ = Me, f : R″ = 2,4,6‐Me3C6H2, g : R″ = SiH(CHMe2)2, h : R″ = SiH(CMe3)2; 2 i : R = R′ = CMe3, R″ = SiH(CMe3)2) were prepared by reaction of the corresponding hydridosilylamines RR′(H)Si–NHR″ 2 a – 2 i with n‐butyllithium in equimolar ratio in n‐hexane. The unknown amines 1 e – 1 i and amides 2 f – 2 i have been characterized spectroscopically. The wave numbers of the Si–H stretching vibrations and 29Si–1H coupling constants of the amides are less than of the analogous amines. This indicates a higher hydride character for the hydrogen atom of the Si–H group in the amide in comparison to the amines. The 29Si‐NMR chemical shifts lie in the amides at higher field than in the amines. The amides 2 a – 2 c and 2 e – 2 g react with chlorotrimethylsilane in THF to give the corresponding N‐silylation products RR′(H)Si–N(SiMe3)R″ ( 3 a – 3 c , 3 e – 3 g ) in good yields. In the reaction of 2 i with chlorotrimethylsilane in molar ratio 1 : 2,33 in THF hydrogen‐chlorine exchange takes place and after hydrolytic work up of the reaction mixture [(Me3C)2(Cl)Si]2NH ( 5 a ) is obtained. The reaction of the amides 2 a – 2 c , 2 f and 2 g with chlorotrimethylsilane in m(p)‐xylene and/or n‐hexane affords mixtures of N‐substitution products RR′(H)Si–N(SiMe3)R″ ( 3 a – 3 c , 3 f , 3 g ) and cyclodisilazanes [RR′Si–NR″]2 ( 6 a – 6 c , 6 f , 6 g ) as the main products. In case of the reaction of 2 h the cyclodisilazane 6 h was obtained only. 2 c – 2 e show a very low reactivity toward chlorotrimetyhlsilane in m‐xylene and toluene resp.. In contrast to Me3SiCl the reactivity of 2 d toward Me3SiOSO2CF3 and Me2(H)SiCl is significant higher. 2 d react with Me3SiOSO2CF3 and Me2(H)SiCl in n‐hexane under N‐silylation to give RR′(H)Si–N(SiMe3)R″ ( 3 d ) and RR′(H)Si–N(SiHMe2)R″ ( 3 d ′) resp. The crystal structures of [Me2Si–NSiMe3]2 ( I ) ( 6 f , 6 g and 6 h ) have been determined.  相似文献   

15.
A series of cationic rare‐earth aryloxide complexes, i.e., [LREOAr']+[B(C6F5)4] (L = CH3C(NAr)CHC(CH3)(NCH(R)CH2PPh2); RE = Y, Lu; Ar' =2,6‐tBu2‐C6H3, 2,6‐(PhCMe2)2‐4‐Me‐C6H2; Ar = 2,6‐iPr2‐C6H3, 2,6‐(Ph2CH)2‐4‐iPr‐C6H2; R = H, CH3, iPr, Ph), were prepared and applied to the Lewis pair polymerization of methyl methacrylate (MMA). The stereoregularity of the resulting PMMA was significantly affected by the R substituent on the pendant arm of the tridentate NNP ligand, and was found to increase with increase in the steric hindrance of R. When using a Ph group as R, the Y complex produced a highly isotactic polymer with an mm value of 95% and a Tg of 54.6 oC. In contrast, the steric hindrance of the Ar and Ar' groups had no effect on the tacticity of the resulting polymer, presumably because these two substituents were situated such that they pointed outward from the cyclic intermediates. Kinetics studies demonstrated that the polymerization was a first‐order process with regard to the monomer concentration prior to catalyst deactivation. End group analysis indicated that the polymerization was accompanied by two possibly competing chain‐termination side reactions that proceeded via intramolecular backbiting cyclization.  相似文献   

16.
Alkane elimination reaction between Ln(CH2SiMe3)3(THF)2 (Ln = Y, Lu) with one equivalent of the amidines with different steric demanding HL ([CyC(N-2,6-iPr2C6H3)2]H (HL1), [CyC(N-2,6-Me2C6H3)2]H (HL2), [PhC(N-2,6-Me2C6H3)2]H (HL3)) in THF afforded a series of mono(amidinate) rare earth metal bis(alkyl) complexes [CyC(N-2,6-iPr2C6H3)2]Ln(CH2SiMe3)2(THF) (Ln = Y (1), Lu (3)), [CyC(N-2,6-Me2C6H3)2]Ln(CH2SiMe3)2(THF)2 (Ln = Y (4), Lu (6)), and [PhC(N-2,6-Me2C6H3)2]Y(CH2SiMe3)2(THF)2 (7) in 75–89% isolated yields. For the early lanthanide metal Nd, THF slurry of NdCl3 was stirred with three equiv of LiCH2SiMe3 in THF, followed by addition of one equiv of the amidines HL1 or HL2 gave an “ate” complex [CyC(N-2,6-iPr2C6H3)2]Nd(CH2SiMe3)2(μ-Cl)Li(THF)3 (2) in 48% yield and a neutral [CyC(N-2,6-Me2C6H3)2]Nd(CH2SiMe3)2(THF)2 (5) in 52% yield, respectively. They were characterized by elemental analysis, FT-IR, NMR spectroscopy (except for 2 and 5 for their strong paramagnetic property). Complexes 2, 3, 4 and 5 were subjected to X-ray single crystal structure determination. These neutral mono(amidinate) rare earth metal bis(alkyl) complexes showed activity towards l-lactide polymerization to give high molecular weight and narrow molecular weight distribution polymers.  相似文献   

17.
The 1‐azonia‐2‐boratanaphthalenes (NH)(BX)C8H6 can be synthesized from 2‐aminostyrene and the dihaloboranes XBHal2 ( 1 ‐ 4 : X = Cl, Br, iPr, tBu). Further derivatives (NH)(BX)C8H6 are obtained from 1 by replacing Cl by alkoxy or alkyl groups [ 5 ‐ 8 : X = OMe, OtBu, Me, (CH2)3NMe2]. The hydrolysis of 1 gives a mixture of the bis(azoniaboratanaphthyl) oxide [(NH)BC8H6]2O ( 9 ) and the hydroxy derivative (NH)[B(OH)]C8H6 ( 10 ). The diboryl oxide 9 crystallizes in the space group C2/c. The lithiation of 4 at the nitrogen atom gives [NLi(tmen)](BtBu)C8H6 ( 11 ), which upon reaction with the diborane(4) B2Cl2(NMe2)2 yields the 1, 2‐bis(azoniaboratanaphthyl)diborane B2[N(BtBu)C8H6]2(NMe2)2 ( 12 ). The 2‐chloro‐1‐methyl‐4‐phenyl derivative (NMe)(BCl)C8H5Ph ( 13 ) of the parent (NH)(BH)C8H6 can be synthesized from the aminoborane BCl2(NMePh) and phenylethyne. Substitution of Cl in 13 gives the derivatives (NMe)(BX)C8H5Ph [ 14 ‐ 20 : X = N(SiMe3)2, Me, Et, iBu, tBu, CH2SiMe3, Ph] and the reaction of 13 with Li2O affords the bis(azoniaboratanaphthyl) oxide [(NMe)BC8H5Ph]2O ( 21 ). The reaction of 16 or 19 with [(MeCN)3Cr(CO)3] yields the complexes [{(NMe)(BX)C8H5Ph}Cr(CO)3] ( 22 , 23 : X = Et, CH2SiMe3), in which the chromium atom is hexahapto bound to the homoarene part of 16 or 19 , respectively. The complex 23 crystallizes in the space group P21/c. Upon reaction of the phenols para‐C6H4R(OH) with the aryldichloroboranes ArBCl2 and subsequent condensation of the products with phenylethyne, the 1‐oxonia‐2‐boratanaphthalenes O(BAr)C8H4RPh with R in position 6 and Ph in position 4 are formed ( 24 ‐ 26 : Ar = Ph, R = H, Me, OMe; 27 ‐ 29 : Ar = C6F5, R = H, Me, OMe). The azoniaboratanaphthalenes 1 ‐ 23 were characterized by NMR methods.  相似文献   

18.
The synthesis of trimethylene-bridged carboxylate-substituted tetraorganodistannoxanes {[Me3SiCH2(RCOO)Sn(CH2)3Sn(OOCR)CH2SiMe3]O} n (1, R = Ph; 2, R = 2,4-Me2C6H3) is reported. Depending on the structure of R, in the solid state these compounds are either dimers (1, n = 2, cis-isomer) with a ladder-type structure or tetramers (2, n = 4) with a double ladder-type structure.  相似文献   

19.
Polymerization of styrene using β‐diketiminate nickel (II) bromide complexes CH{C(R)NAr}2NiBr (R = CH3, Ar = 2,6‐iPr2C6H3, 1 ; R = CH3, Ar = 2,6‐Me2C6H3, 2 ; R = CF3, Ar = 2,6‐iPr2C6H3, 3 ; R = CF3, Ar = 2,6‐Me2C6H3, 4 ) in the presence of methylaluminoxane was studied. Compound 3 is the most active styrene polymerization catalyst of all the nickel complexes tested. The activity of these catalysts increases with increases in steric bulk of the substituents on the aryl rings. The electronic nature of the ligand backbone also affects the activity. Weight‐average molecular weight of the prepared polystyrene ranges from 21 000 to 72 000, with polydispersity indexes of 1.95–2.78. The microstructure of the obtained products is atactic polystyrenes from NMR analyses. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
The complexes [cis-Rh(SC6F5)(PPh3)2(L)] (L = py, 3-Mepy, isoquin, N-Melm; py = pyridine, 3-Mepy = 3-methylpyridine, isoquin = isoquinoline, N-Melm = N-methylimidazole) readily undergo oxidative addition of HR (R = H, SC6F5, C2Ph) to give [RhH(R)(SC6F5)(PPh3)n(L)3−n] (n = 1, 2) whereas the complexes [cis-Rh(SC6F5)(PPh3)2(L′)] (L′ = 2-Mepy, 2,6-Me2py, quin; 2-Mepy = 2-methylpyridine; 2,6-Me2py = 2,6-dimethylpyridine, quin = quinoline) react only where R = C2Ph. Where conditions favour the formation of [RhH(R)(SC6F5)(PPh3)n(L′)3−n] reductive elimination of H2 (R = H) or C6F5SH (R = SC6F5, C2Ph) occurs.  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