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.
Surface organometallic chemistry (SOMC) on silica materials is a prominent approach for the generation of highly active heterogenized polymerization catalysts. Despite advanced methods of characterization, the elucidation of the catalytically active surface species remains a challenging task. Alkylated rare‐earth metal siloxide complexes can be regarded as molecular models of respective covalently bonded alkylated surface species, primarily used for 1,3‐diene polymerization. Here, we performed both salt metathesis reactions of [Y(MMe4)3] (M = Al, Ga) with [K{OSi(OtBu)3}] and alkylation reactions of [Y{OSi(OtBu)3}3]2 with AlMe3. The obtained complexes [Y(CH3)[(AlMe2){OSi(OtBu)3}2](AlMe4)]2, [Y(CH3)[(AlMe2){OSi(OtBu)3}2]‐{OSi(OtBu)3}], [Y{OSi(OtBu)3}3(μ‐Me)Y(μ‐Me)2Y{OSi(OtBu)3}2(AlMe4)], and [Y(CH3)(GaMe4){OSi(OtBu)3}]2 represent rare examples of organoyttrium species with terminal methyl groups. The formation and purity of the mixed methyl/siloxy yttrium complexes could be enhanced by treating [Y(MMe4)3] with [K(MMe2){OSi(OtBu)3}2]n (M=Al: n=2; M=Ga: n=∞). Complexes [K(MMe2){OSi(OtBu)3}2]n were obtained by addition of [K{OSi(OtBu)3}] to [Me2M{OSi(OtBu)3}]2. Deeper insight into the fluxional behavior of the mixed methyl/siloxy yttrium complexes in solution was gained by 1H and 13C NMR spectroscopic studies at variable temperature and 1H–89Y HSQC NMR spectroscopy.  相似文献   

2.
The N‐heterocyclic carbene (NHC) adducts Zn(CpR)2(NHC)] (CpR=C5HMe4, C5H4SiMe3; NHC=ItBu, IDipp (Dipp=2,6‐diisopropylphenyl), IMes (Mes=mesityl), SIMes) were prepared and shown to be active catalysts for the hydrogenation of imines, whereas decamethylzincocene [ZnCp*2] is highly active for the hydrogenation of ketones in the presence of noncoordinating NHCs. The abnormal carbene complex [Zn(OCHPh2)2(aItBu)]2 was formed from spontaneous rearrangement of the ItBu ligand during incomplete hydrogenation of benzophenone. Two isolated ZnI adducts [Zn2Cp*2(NHC)] (NHC=ItBu, SIMes) are presented and characterized as weak adducts on the basis of 13C NMR spectroscopic and X‐ray diffraction experiments. A mechanistic proposal for the reduction of [ZnCp*2] with H2 to give [Zn2Cp*2] is discussed.  相似文献   

3.
A general regioselective rhodium‐catalyzed head‐to‐tail dimerization of terminal alkynes is presented. The presence of a pyridine ligand (py) in a Rh–N‐heterocyclic‐carbene (NHC) catalytic system not only dramatically switches the chemoselectivity from alkyne cyclotrimerization to dimerization but also enhances the catalytic activity. Several intermediates have been detected in the catalytic process, including the π‐alkyne‐coordinated RhI species [RhCl(NHC)(η2‐HC?CCH2Ph)(py)] ( 3 ) and [RhCl(NHC){η2‐C(tBu)?C(E)CH?CHtBu}(py)] ( 4 ) and the RhIII–hydride–alkynyl species [RhClH{? C?CSi(Me)3}(IPr)(py)2] ( 5 ). Computational DFT studies reveal an operational mechanism consisting of sequential alkyne C? H oxidative addition, alkyne insertion, and reductive elimination. A 2,1‐hydrometalation of the alkyne is the more favorable pathway in accordance with a head‐to‐tail selectivity.  相似文献   

4.
The reactivity towards AlMe3 of discrete cationic ansa‐zirconocenes 2 a,b that are ubiquitously used in isoselective propylene polymerization and based on [{Ph(H)C(3,6‐tBu2‐Flu)(3‐tBu‐5‐Et‐Cp)}ZrMe2)] {Cp‐Flu} and rac‐[{Me2Si‐(2‐Me‐4‐Ph‐Ind)2}ZrMe2] {SBI} was scrutinized. The first example of a structurally characterized Group 4 metallocene AlMe3 adduct ( 3 b ) is reported. In the presence of excess AlMe3, the {SBI}‐based AlMe3 adduct 3 b undergoes a slow decomposition via C? H activation in a bridging methyl unit to yield a new species ( 4 b ) with a trimetallic {Zr(μ‐CH2)(μ‐Me)AlMe(μ‐Me)AlMe2} core. EXSY NMR data for the process 2 b ? 3 b → 4 b suggest very rapid and reversible binding of an additional AlMe3 molecule onto AlMe3 adduct 3 b . The resulting heterotrimetallic species intermediates exchange of methyl groups between different metal centers and slowly undergoes the C? H activation reaction towards 4 b .  相似文献   

5.
The reaction of monomeric [(TptBu,Me)LuMe2] (TptBu,Me=tris(3‐Me‐5‐tBu‐pyrazolyl)borate) with primary aliphatic amines H2NR (R=tBu, Ad=adamantyl) led to lutetium methyl primary amide complexes [(TptBu,Me)LuMe(NHR)], the solid‐state structures of which were determined by XRD analyses. The mixed methyl/tetramethylaluminate compounds [(TptBu,Me)LnMe({μ2‐Me}AlMe3)] (Ln=Y, Ho) reacted selectively and in high yield with H2NR, according to methane elimination, to afford heterobimetallic complexes: [(TptBu,Me)Ln({μ2‐Me}AlMe2)(μ2‐NR)] (Ln=Y, Ho). X‐ray structure analyses revealed that the monomeric alkylaluminum‐supported imide complexes were isostructural, featuring bridging methyl and imido ligands. Deeper insight into the fluxional behavior in solution was gained by 1H and 13C NMR spectroscopic studies at variable temperatures and 1H–89Y HSQC NMR spectroscopy. Treatment of [(TptBu,Me)LnMe(AlMe4)] with H2NtBu gave dimethyl compounds [(TptBu,Me)LnMe2] as minor side products for the mid‐sized metals yttrium and holmium and in high yield for the smaller lutetium. Preparative‐scale amounts of complexes [(TptBu,Me)LnMe2] (Ln=Y, Ho, Lu) were made accessible through aluminate cleavage of [(TptBu,Me)LnMe(AlMe4)] with N,N,N′,N′‐tetramethylethylenediamine (tmeda). The solid‐state structures of [(TptBu,Me)HoMe(AlMe4)] and [(TptBu,Me)HoMe2] were analyzed by XRD.  相似文献   

6.
Comparative Structural Studies on 4‐Dimethylaminopyridine‐Adducts Lewis acid‐base adducts of the type dmap—MMe3 (M = Al 1 , Ga 2 , In 3 , Tl 4 ) as well as dmap—AlCl3 ( 6 ) and dmap—Al(t‐Bu)3 ( 7 ) were synthesized by reaction of MR3 with 4‐dimethylamino‐pyridine (dmap) whereas dmap—AlH3 ( 5 ) was obtained from AlH3·Et2O. 1 — 7 were characterized by means of NMR (1H, 13C{1H}) and mass spectrometry and elemental analysis. In addition, their solid state structures were determined by single crystal X‐ray diffraction studies. A comparison of the structural parameters reveales the influence of both electronic (Lewis acidity of the group 13 atom) and steric interactions on the structure and stability of as prepared Lewis acid‐base adducts.  相似文献   

7.
The mononuclear N‐heterocyclic carbene (NHC) copper alkoxide complexes [(6‐NHC)CuOtBu] (6‐NHC=6‐MesDAC ( 1 ), 6‐Mes ( 2 )) have been prepared by addition of the free carbenes to the tetrameric tert‐butoxide precursor [Cu(OtBu)]4, or by protonolysis of [(6‐NHC)CuMes] (6‐NHC=6‐MesDAC ( 3 ), 6‐Mes ( 4 )) with tBuOH. In contrast to the relatively stable diaminocarbene complex 2 , the diamidocarbene derivative 1 proved susceptible to both thermal and hydrolytic ring‐opening reactions, the latter affording [(6‐MesDAC)Cu(OC(O)CMe2C(O)N(H)Mes)(CNMes)] ( 6 ). The intermediacy of [(6‐MesDAC)Cu(OH)] in this reaction was supported by the generation of Cu2O as an additional product. Attempts to generate an isolable copper hydride complex of the type [(6‐MesDAC)CuH] by reaction of 1 with Et3SiH resulted instead in migratory insertion to generate [(6‐MesDAC‐H)Cu(P(p‐tolyl)3)] ( 9 ) upon trapping by P(p‐tolyl)3. Migratory insertion was also observed during attempts to prepare [(6‐Mes)CuH], with [(6‐Mes‐H)Cu(6‐Mes)] ( 10 ) isolated, following a reaction that was significantly slower than in the 6‐MesDAC case. The longer lifetime of [(6‐Mes)CuH] allowed it to be trapped stoichiometrically by alkyne, and also employed in the catalytic semi‐reduction of alkynes and hydrosilylation of ketones.  相似文献   

8.
The half‐open rare‐earth‐metal aluminabenzene complexes [(1‐Me‐3,5‐tBu2‐C5H3Al)(μ‐Me)Ln(2,4‐dtbp)] (Ln=Y, Lu) are accessible via a salt metathesis reaction employing Ln(AlMe4)3 and K(2,4‐dtbp). Treatment of the yttrium complex with B(C6F5)3 and tBuCCH gives access to the pentafluorophenylalane complex [{1‐(C6F5)‐3,5‐tBu2‐C5H3Al}{μ‐C6F5}Y{2,4‐dtbp}] and the mixed vinyl acetylide complex [(2,4‐dtbp)Y(μ‐η13‐2,4‐tBu2‐C5H4)(μ‐CCtBu)AlMe2], respectively.  相似文献   

9.
Syntheses and Reactions of Aluminium Alkoxide Compounds Al(OcHex)3 ( 1 ) can be synthesized by the reaction of Al with cyclohexanol under evolving of H2 in boiling xylene. [Li{Al(OCH2Ph)4}] ( 2 ) was obtained by treatment of PhCH2OH with a 1 M solution of LiAlH4 in THF. [{(THF)Li}2{Al(OtBu)4}Cl] ( 3 ) is the result of the reaction of four equivalents of LiOtBu on AlCl3 in THF. 3 is the educt for the reactions with the Lewis‐acids InCl3 and FeCl3 in THF leading to the metalates [{(THF)2Li}2{Al(OtBu)4}] · [MCl4] [M = In ( 4 ), Fe ( 5 )]. The attempt to react InCl3 with four equivalents of LiOtBu leads to only one isolated and characterized product, the complex [Li4(OtBu)3(THF)3Cl]2 · THF ( 6 · THF), which can also be synthesized by the treatment of LiCl with three equivalents of LiOtBu in THF. 1–6 · THF were characterized by NMR, IR and MS techniques as well as by X‐ray structure determinations. According to them, 1 , which is tetrameric in solution, is the first structurally characterized example of the proposed trimer form of aluminium alkoxides [ROAl{Al(OR)4}2] with a central trigonal bipyramidal coordinated Al atom. 2 forms a coordination polymer with a distorted tetrahedral coordination sphere of Li and Al, running along [100]. The trinuclear structure skeleton [{(THF)2Li}2{Al(OtBu)4}]+ is still present in the isotypical metalates 4 and 5 . The counter ions [MCl4] possess nearly Td symmetry. The remarkable structural motif of 6 · THF are two heterocubanes [Li4(OtBu)3(THF)3Cl] dimerized by Li–Cl bonds.  相似文献   

10.
Herein the sodium alkylmagnesium amide [Na4Mg2(TMP)6(nBu)2] (TMP=2,2,6,6‐tetramethylpiperidide), a template base as its deprotonating action is dictated primarily by its 12 atom ring structure, is studied with the common N‐heterocyclic carbene (NHC) IPr [1,3‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene]. Remarkably, magnesiation of IPr occurs at the para‐position of an aryl substituent, sodiation occurs at the abnormal C4 position, and a dative bond occurs between normal C2 and sodium, all within a 20 atom ring structure accommodating two IPr2?. Studies with different K/Mg and Na/Mg bimetallic bases led to two other magnesiated NHC structures containing two or three IPr? monoanions bound to Mg through abnormal C4 sites. Synergistic in that magnesiation can only work through alkali‐metal mediation, these reactions add magnesium to the small cartel of metals capable of directly metalating a NHC.  相似文献   

11.
An in depth study of the reactivity of an N‐heterocyclic carbene (NHC)‐stabilized silylene monohydride with alkynes is reported. The reaction of silylene monohydride 1 , tBu3Si(H)Si←NHC, with diphenylacetylene afforded silole 2 , tBu3Si(H)Si(C4Ph4). The density functional theory (DFT) calculations for the reaction mechanism of the [2+2+1] cycloaddition revealed that the NHC played a major part stabilizing zwitterionic transition states and intermediates to assist the cyclization pathway. A significantly different outcome was observed, when silylene monohydride 1 was treated with phenylacetylene, which gave rise to supersilyl substituted 1‐alkenyl‐1‐alkynylsilane 3 , tBu3Si(H)Si(CH?CHPh)(C?CPh). Mechanistic investigations using an isotope labelling technique and DFT calculations suggest that this reaction occurs through a similar zwitterionic intermediate and subsequent hydrogen abstraction from a second molecule of phenylacetylene.  相似文献   

12.
Lithium and Cesium Alkoxometalates The aluminium alkoxide, Al(OCH2Ph)3 ( 1 ), can be obtained from a direct synthesis of Al and PhCH2OH under HgCl2 catalysis. The formation of the metalate [{(Diglyme)Li}{Al(OtBu)4}] ( 2 ) from LiAlH4 and tBuOH in THF under evolution of hydrogen takes place, if the reaction product is heated under reflux with additional tBuOH in diglyme. The nucleophilic attack of F ions leads during the treatment of CsF on a THF solution of Al(OcHex)3 after ligand redistribution to the coordination polymer [{Cs(THF)2}{Cs(THF)}{Al(OcHex)4}2]n ([3]n). 1 , 2 , and 3 were characterized by NMR, IR and MS techniques as well as by crystal structure analyses. According to them 1 is present as tetramer in solution and the solid state. The central structural motif of the metalate 2 is a heteronuclear and planar LiO2Al four‐membered ring with a penta‐coordinated Li+ ion. In the chainlike coordination polymer [ 3 ]n Cs+ ions with coordination number five and six occupy alternating positions.  相似文献   

13.
The synthesis and reactivity of a silyliumylidene cation stabilized by an amidinate ligand and 4‐dimethylaminopyridine (DMAP) are described. The reaction of the amidinate silicon(I) dimer [ L Si:]2 ( 1 ; L =PhC(NtBu)2) with one equivalent of N‐trimethylsilyl‐4‐dimethylaminopyridinium triflate [4‐NMe2C5H4NSiMe3]OTf and two equivalents of DMAP in THF afforded [ L Si(DMAP)]OTf ( 2 ). The ambiphilic character of 2 is demonstrated from its reactivity. Treatment of 2 with 1 in THF afforded the disilylenylsilylium triflate [ L′ 2( L )Si]OTf ( 3 ; L′ = L Si:) with the displacement of DMAP. The reaction of 2 with [K{HB(iBu)3}] and elemental sulfur in THF afforded the silylsilylene [ L SiSi(H){(NtBu)2C(H)Ph}] ( 4 ) and the base‐stabilized silanethionium triflate [ L Si(S)DMAP]OTf ( 5 ), respectively. Compounds 2 , 3 , and 5 have been characterized by X‐ray crystallography.  相似文献   

14.
This study reports the use of the trisalkylgallium GaR3 (R=CH2SiMe3), containing sterically demanding monosilyl groups, as an effective Lewis‐acid component for frustrated Lewis pair activation of carbonyl compounds, when combined with the bulky N‐heterocyclic carbene 1,3‐bis(tert‐butyl)imidazol‐2‐ylidene (ItBu) or 1,3‐bis(tert‐butyl)imidazolin‐2‐ylidene (SItBu). The reduction of aldehydes can be achieved by insertion into the C=O functionality at the C2 (so‐called normal) position of the carbene affording zwitterionic products [ItBuCH2OGaR3] ( 1 ) or [ItBuCH(p‐Br‐C6H4)OGaR3] ( 2 ), or alternatively, at its abnormal (C4) site yielding [aItBuCH(p‐Br‐C6H4)OGaR3] ( 3 ). As evidence of the cooperative behaviour of both components, ItBu and GaR3, neither of them alone are able to activate any of the carbonyl‐containing substrates included in this study NMR spectroscopic studies of the new compounds point to complex equilibria involving the formation of kinetic and thermodynamic species as implicated through DFT calculations. Extension to ketones proved successful for electrophilic α,α,α‐trifluoroacetophenone, yielding [aItBuC(Ph)(CF3)OGaR3] ( 7 ). However, in the case of ketones and nitriles bearing acidic hydrogen atoms, C?H bond activation takes place preferentially, affording novel imidazolium gallate salts such as [{ItBuH}+{(p‐I‐C6H4)C(CH2)OGaR3}?] ( 8 ) or [{ItBuH}+{Ph2C=C=NGaR3}?] ( 12 ).  相似文献   

15.
Metalat Ions [Al(OR)4] as Chelating Ligands for Transition Metal Cations Waterfree CoCl2 can be reacted with [{Li(Diglyme)}{Al(OtBu)4}] in THF to the complex [Li(THF)4][{CoCl2}{Al(OtBu)4}]. Addition of diglyme to the reaction mixtures gives the blue compound [Li(diglyme)2][{CoCl2}{Al(OtBu)4}] ( 1 ). According to this procedure the FeII complex [Li(Diglyme)2][{FeCl2}2{Al(OtBu)4}] ( 2 ) was formed by treatment of FeCl2 with Li[Al(OtBu)4]. [{Li(diglyme)}{Al(OtBu)4}] in THF/diglyme can be used as alkoxide transfer reagent on TiCl4 to give the neutral complex [TiCl2(OtBu)2(diglyme)] ( 3 ). The sky‐blue salt [Li(THF)4]2[{CoCl2}3{Al(OCH2Ph)4}2] ( 4 ) was obtained by reaction of Li[Al(OCH2Ph)4] with CoCl2 in THF. By treatment of 4 with diglyme ligand redistribution was observed giving the sky‐blue compound [Li(Diglyme)2]2[{CoCl2}3{Al(OCH2Ph)4}2] ( 5 ) and the violet salt [Li(Diglyme)2]2[Co2Cl5(OCH2Ph)] ( 6 ). A similar salt can be synthesized also directly from Li[Al(OtBu)4] and CoCl2 in diglyme to give [Li(Diglyme)2]2[Co2Cl5(OtBu)] ( 7 ). 1 — 7 were characterized by IR spectroscopy, partly by mass spectrometry and X‐ray analyses. UV‐VIS spectra were recorded from 1 and 5 . According to the X‐ray analyses the MII ions as well as the AlIII ions are coordinated distorted tedrahedrally. In 1 , 2 , 4 und 5 the unit [Al(OR)4] acts a chelating ligand as desired.  相似文献   

16.
Ansa‐zirconocene diamide complex rac‐(EBI)Zr(NMe2)2 [rac‐1, EBI = ethylene‐1,2‐bis(1‐indenyl)] reacted with AlR3 (R = Me, Et, iBu) or Al(iBu2)H and then with [CPh3][B(C6F5)4] (2) in toluene in order to perform propylene polymerization by cationic alkylzirconium species, which are in situ generated during polymerization. Through the sequential NMR‐scale reactions of rac‐1 with AlR3 or Al(iBu2)H and then with 2, rac‐1 was demonstrated to be transformed to the active alkyzirconium cations via alkylated intermediates of rac‐1. The cationic species generated by using AlMe3, AlEt3, and Al(iBu2)H as alkylating reagents tend to become heterodinuclear complex; however, those by using bulky Al(iBu)3 become base‐free [rac‐(EBI)Zr(iBu)]+ cations. The activity of propylene polymerization by rac‐1/AlR3/2 catalyst was deeply influenced by various parameters such as the amount and the type of AlR3, metallocene concentration, [Al]/[2] ratio, and polymerization temperature. Generally the catalytic systems using bulky alkylaluminum like Al(iBu)3 and Al(iBu)2H show higher activity but lower stereoregularity than those using less bulky AlMe3 and AlEt3. The alkylating reagent Al(iBu)3 is not a transfer agent as good as AlMe3 or AlEt3. The polymerization activities show maximum around [Al]/[2] ratio of 1.0 and increase monotonously with polymerization temperature. The overall activation energy of both rac‐1/Al(iBu)3/2 and rac‐1/Al(iBu)2H catalysts is 6.0 kcal/mol. As the polymerization temperature increases, the stereoregularity of the resulting polymer decreases markedly, which is demonstrated by the decrease of [mmmm] pentad value and by the increase of the amount of polymer soluble in low boiling solvent. The physical properties of polymers produced in this study were investigated by using 13C‐NMR, differential scanning calorimetry (DSC), viscometry, and gel permeation chromatography (GPC). © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1523–1539, 1999  相似文献   

17.
A series of new, easily activated NHC–PdII precatalysts featuring a trans‐oriented morpholine ligand were prepared and evaluated for activity in carbon‐sulfur cross‐coupling chemistry. [(IPent)PdCl2(morpholine)] (IPent=1,3‐bis(2,6‐di(3‐pentyl)phenyl)imidazol‐2‐ylidene) was identified as the most active precatalyst and was shown to effectively couple a wide variety of deactivated aryl halides with both aryl and alkyl thiols at or near ambient temperature, without the need for additives, external activators, or pre‐activation steps. Mechanistic studies revealed that, in contrast to other common NHC–PdII precatalysts, these complexes are rapidly reduced to the active NHC–Pd0 species at ambient temperature in the presence of KOtBu, thus avoiding the formation of deleterious off‐cycle PdII–thiolate resting states.  相似文献   

18.
A copper/N‐heterocyclic carbene (NHC) catalyzed oxidative vicinal diphosphination of styrenes with diphenyl(trimethylsilyl)phosphine proceeds in the presence of LiOtBu and a pyridine N‐oxide/MnO2 combined oxidant to deliver the corresponding 1,2‐bis(diphenylphosphino)ethanes (DPPEs) in good yields. The present copper catalysis can provide access to the DPPE‐type ligands directly from the relatively simple alkenes.  相似文献   

19.
Ansa‐zirconocene diamide complex rac‐Me2Si(CMB)2Zr(NMe2)2 (rac‐1, CMB = 1‐C5H2‐2‐Me‐4‐tBu) reacts with AlR3 (R = Me, Et, i‐Bu) and then with [CPh3]+[B(C6F5)4] (2) in toluene in order to in situ generate cationic alkylzirconium species. In the sequential NMR‐scale reactions of rac‐1 with various amount of AlMe3 and 2, rac‐1 transforms first to rac‐Me2Si(CMB)2Zr(Me)(NMe2) (rac‐3) and rac‐Me2Si(CMB)2ZrMe2 (rac‐4) by the reaction with AlMe3, and then to [rac‐Me2Si(CMB)2ZrMe]+ (5+) cation by the reaction of the resulting mixtures with 2. The activities of propylene polymerizations by rac‐1/Al(i‐Bu)3/2 system are dependent on the type and concentration of AlR3, resulting in the order of activity: rac‐1/Al(i‐Bu)3/2 > rac‐1/AlEt3/2 > rac‐1/MAO ≫ rac‐1/AlMe3/2 system. The bulkier isobutyl substituents make inactive catalytic species sterically unfavorable and give rise to more separated ion pairs so that the monomers can easily access to the active sites. The dependence of the maximum rate (Rp, max) on polymerization temperature (Tp) obtained by rac‐1/Al(i‐Bu)3/2 system follows Arrhenius relation, and the overall activation energy corresponds to 0.34 kcal/mol. The molecular weight (MW) of the resulting isotactic polypropylene (iPP) is not sensitive to Al(i‐Bu)3 concentration. The analysis of regiochemical errors of iPP shows that the chain transfer to Al(i‐Bu)3 is a minor chain termination. The 1,3‐addition of propylene monomer is the main source of regiochemical sequence and the [mr] sequence is negligible, as a result the meso pentad ([mmmm]) values of iPPs are very high ([mmmm] > 94%). These results can explain the fact that rac‐1/Al(i‐Bu)3/2 system keeps high activity over a wide range of [Al(i‐Bu)3]/[Zr] ratio between 32 and 3,260. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1071–1082, 1999  相似文献   

20.
Dilithiated N,N′‐dimethyl‐piperazine, LiCH2N(CH2CH2)2 NCH2Li ( 2 ) was prepared by transmetallation of N,N′‐bis(trimethylstannylmethyl)‐piperazine ( 1 ) with nBuLi and was isolated as a highly pyrophoric yellowish powder in high yield. Compound 2 was characterized by elemental analysis and was reacted as difunctional aminomethylating reagent with dialkyl‐earth metal chlorides, R2MCl (M = Al, Ga; R = Me, tBu) which resulted in the formation of spirocyclic adducts of N,N′‐bis(dialkylmetallamethyl)‐piperazine and unreacted dialkylmetal chlorides, [(Me2AlCl)Me2AlCH2N(CH2CH2)2NCH2AlMe2(ClAlMe2)] ( 3 ) and [(tBu2GaCl)tBu2GaCH2N(CH2CH2)2NCH2GatBu2(ClGatBu2)] ( 4 ) with five‐membered rings. Compounds 1 , 3 and 4 were identified by NMR‐spectroscopy (1H, 13C, 119Sn for 1 , 27Al for 3 ), mass spectra (EI, for 1 ) and by crystal structure determinations.  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