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.
The AlI compound NacNacAl ( 1 , NacNac = [ArNC(Me)CHC(Me)NAr], Ar = 2,6-iPr2C6H3) serves as a template for the chemoselective coupling between carbonyls (benzophenone, fenchone, isophorone, p-tolyl benzoate, N,N-dimethylbenzamide, (1-phenylethylidene)aniline) and pyridine. With the CH-acidic ketone (1R)-(+) camphor, the reaction affords a hydrido alkoxide compound of Al, formed as the result of enolization, whereas an enolizable imine, (1-phenylethylidene)aniline, and the bulky ketone isophorone, still chemoselectively couple with pyridine. In contrast, reaction with the ester p-tolyl benzoate results in cleavage of the ester bond together with replacement of the alkoxy group by a hydrogen atom of the pyridine moiety. This study demonstrates that for carbonyl substrates featuring phenyl substituents, the reaction proceeds via intermediate formation of η2(C,X)-coordinated (X = O, N) carbonyl adducts, whereas the reaction of 1 with (R)-(−)-fenchone in the absence of pyridine leads to CH activation in the pendant isopropyl group of the Ar substituent of the NacNac ligand.  相似文献   

2.
The aluminum(I) compound NacNacAl (NacNac=[ArNC(Me)CHC(Me)NAr], Ar=2,6-iPr2C6H3, 1 ) shows diverse and substrate-controlled reactivity in reactions with N-heterocycles. 4-Dimethylaminopyridine (DMAP), a basic substrate in which the 4-position is blocked, induces rearrangement of NacNacAl by shifting a hydrogen atom from the methyl group of the NacNac backbone to the aluminum center. In contrast, C−H activation of the methyl group of 4-picoline takes place to produce a species with a reactive terminal methylene. Reaction of 1 with 3,5-lutidine results in the first example of an uncatalyzed, room-temperature cleavage of an sp2 C−H bond (in the 4-position) by an AlI species. Another reactivity mode was observed for quinoline, which undergoes 2,2′-coupling. Finally, the reaction of 1 with phthalazine produces the product of N−N bond cleavage.  相似文献   

3.
The aluminum(I) compound NacNacAl (NacNac=[ArNC(Me)CHC(Me)NAr]?, Ar=2,6‐iPr2C6H3, 1 ) shows diverse and substrate‐controlled reactivity in reactions with N‐heterocycles. 4‐Dimethylaminopyridine (DMAP), a basic substrate in which the 4‐position is blocked, induces rearrangement of NacNacAl by shifting a hydrogen atom from the methyl group of the NacNac backbone to the aluminum center. In contrast, C?H activation of the methyl group of 4‐picoline takes place to produce a species with a reactive terminal methylene. Reaction of 1 with 3,5‐lutidine results in the first example of an uncatalyzed, room‐temperature cleavage of an sp2 C?H bond (in the 4‐position) by an AlI species. Another reactivity mode was observed for quinoline, which undergoes 2,2′‐coupling. Finally, the reaction of 1 with phthalazine produces the product of N?N bond cleavage.  相似文献   

4.
The treatment of cyclic thioureas with the aluminum(I) compound NacNacAl ( 1 ; NacNac=[ArNC(Me)CHC(Me)NAr]?, Ar=2,6‐Pri2C6H3) resulted in oxidative cleavage of the C=S bond and the formation of 3 and 5 , the first monomeric aluminum complexes with an Al=S double bond stabilized by N‐heterocyclic carbenes. Compound 1 also reacted with triphenylphosphine sulfide in a similar manner, which resulted in cleavage of the P=S bond and production of the adduct [NacNacAl=S(S=PPh3)] ( 8 ). The Al=S double bond in 3 can react with phenyl isothiocyanate to furnish the cycloaddition product 9 and zwitterion 10 as a result of coupling between the liberated carbene and PhN=C=S. All novel complexes were characterized by multinuclear NMR spectroscopy, and the structures of 5 , 9 , and 10 were confirmed by X‐ray diffraction analysis. The nature of the Al=S bond in 5 was also probed by DFT calculations.  相似文献   

5.
Three new iron(II)‐benzilate complexes [(N4Py)FeII(benzilate)]ClO4 ( 1 ), [(N4PyMe2)FeII(benzilate)]ClO4 ( 2 ) and [(N4PyMe4)FeII(benzilate)]ClO4 ( 3 ) of neutral pentadentate nitrogen donor ligands have been isolated and characterized to study their dioxygen reactivity. Single‐crystal X‐ray structures reveal a mononuclear six‐coordinate iron(II) center in each case, where benzilate binds to the iron center in monodentate mode via one carboxylate oxygen. Introduction of methyl groups in the 6‐positions of the pyridine rings makes the N4PyMe2 and N4PyMe4 ligand fields weaker compared to that of the parent N4Py ligand. All the complexes ( 1 – 3 ) react with dioxygen to decarboxylate the coordinated benzilate to benzophenone quantitatively. The decarboxylation is faster for the complex of the more sterically hindered ligand and follows the order 3 > 2 > 1 . The complexes display oxygen atom transfer reactivity to thioanisole and also exhibit hydrogen atom transfer reactions with substrates containing weak C?H bonds. Based on interception studies with external substrates, labelling experiments and Hammett analysis, a nucleophilic iron(II)‐hydroperoxo species is proposed to form upon two‐electron reductive activation of dioxygen by each iron(II)‐benzilate complex. The nucleophilic oxidants are converted to the corresponding electrophilic iron(IV)‐oxo oxidant upon treatment with a protic acid. The high‐spin iron(II)‐benzilate complex with the weakest ligand field results in the formation of a more reactive iron‐oxygen oxidant.  相似文献   

6.
The reactivity of an aluminium(I) diketiminate compound NacNacAl (NacNac is [ArNC(Me)CHC(Me)NAr]?, where Ar is 2,6-diisopropylphenyl) towards arenes has been systematically explored. Heating NacNacAl in benzene results in a fragmentation of the NacNac moiety due to cleavage of the CN bond, while anthracene adds to the main group carbenoid in a [4+1] fashion. Reactions with phenanthrene, triphenylene and fluoranthene demonstrate a reversible [4+1] addition process.  相似文献   

7.
Described is the development of a new class of bis(cyclometalated) ruthenium(II) catalyst precursors for C? C coupling reactions between alkene and alkyne substrates. The complex [(cod)Ru(3‐methallyl)2] reacts with benzophenone imine or benzophenone in a 1:2 ratio to form bis(cyclometalated) ruthenium(II) complexes ( 1 ). The imine‐ligated complex 1 a promoted room‐temperature coupling between acrylic esters and amides with internal alkynes to form 1,3‐diene products. A proposed catalytic cycle involves C? C bond formation by oxidative cyclization, β‐hydride elimination, and C? H bond reductive elimination. This RuII/RuIV pathway is consistent with the observed catalytic reactivity of 1 a for mild tail‐to‐tail methyl acrylate dimerization and for cyclobutene formation by [2+2] norbornene/alkyne cycloaddition.  相似文献   

8.
The stable cationic iridacyclopentenylidene [TpMe2Ir(?CHC(Me)?C(Me)C H2(NCMe)]PF6 ( A ; TpMe2=hydrotris(3,5‐dimethylpyrazolyl)borate) has been obtained by α‐hydride abstraction from the iridacyclopent‐2‐ene [TpMe2Ir(CH2C(Me)?C(Me)C H2)(NCMe)]. Complex A exhibits Brønsted–Lowry acidity at the Ir? CH2 and proximal (relative to Ir? CH2) methyl sites. The coordination of an extra molecule of acetonitrile to the iridium center initiates the reversible isomerization of the chelating carbon chain of A to the monodentate butadienyl ligand of complex [TpMe2Ir(CH?C(Me)C(Me)?CH2)(NCMe)2]PF6, which is capable to engage in a water‐promoted C? C coupling with the MeCN co‐ligands. The product is an aesthetically appealing bicyclic structure that resembles the hydrocarbon barrelene.  相似文献   

9.
The stable cationic iridacyclopentenylidene [TpMe2Ir(CHC(Me)C(Me)C H2(NCMe)]PF6 ( A ; TpMe2=hydrotris(3,5‐dimethylpyrazolyl)borate) has been obtained by α‐hydride abstraction from the iridacyclopent‐2‐ene [TpMe2Ir(CH2C(Me)C(Me)C H2)(NCMe)]. Complex A exhibits Brønsted–Lowry acidity at the Ir CH2 and proximal (relative to Ir CH2) methyl sites. The coordination of an extra molecule of acetonitrile to the iridium center initiates the reversible isomerization of the chelating carbon chain of A to the monodentate butadienyl ligand of complex [TpMe2Ir(CHC(Me)C(Me)CH2)(NCMe)2]PF6, which is capable to engage in a water‐promoted C C coupling with the MeCN co‐ligands. The product is an aesthetically appealing bicyclic structure that resembles the hydrocarbon barrelene.  相似文献   

10.
The reaction of diazo compounds with alkenes catalysed by complex [RuCl(cod)(Cp)] (cod=1,5‐cyclooctadiene, Cp=cyclopentadienyl) has been studied. The catalytic cycle involves in the first step the decomposition of the diazo derivative to afford the reactive [RuCl(Cp){?C(R1)R2}] intermediate and a mechanism is proposed for this step based on a kinetic study of the simple coupling reaction of ethyl diazoacetate. The evolution of the Ru–carbene intermediate in the presence of alkenes depends on the nature of the substituents at both the diazo N2?C(R1)R2 (R1, R2=Ph, H; Ph, CO2Me; Ph, Ph; C(R1)R2=fluorene) and the olefin substrates R3(H)C?C(H)R4 (R3, R4=CO2Et, CO2Et; Ph, Ph; Ph, Me; Ph, H; Me, Br; Me, CN; Ph, CN; H, CN; CN, CN). A remarkable reactivity of the complex was recorded, especially towards unstable aryldiazo compounds and electron‐poor olefins. The results obtained indicate that either cyclopropanation or metathesis products can be formed: the first products are favoured by the presence of a cyano substituent at the double bond and the second ones by a phenyl.  相似文献   

11.
Modifying the β‐diketimine ligand LH 1 (LH=[ArN?C(Me)? CH?C(Me)? NHAr], Ar=2,6‐iPr2C6H3) through replacement of the proton in 3‐position by a benzyl group (Bz) leads to the new BzLH ligand 2, which could be isolated in 77 % yield. According to 1H NMR spectroscopy, 2 is a mixture of the bis(imino) form [(ArN?C(Me)]2CH(Bz) 2a and its tautomer [ArN?C(Me)? C(Bz)?C(Me)NHAr] 2b. Nevertheless, lithiation of the mixture of 2a and 2b affords solely the N‐lithiated β‐diketiminate [ArN?C(Me)? C(Bz)?C(Me)? NLiAr], BzLLi 3. The latter reacts readily with GeCl2?dioxane to form the chlorogermylene BzLGeCl 4, which serves as a precursor for a new zwitterionic germylene by dehydrochlorination with LiN(SiMe3)2. This reaction leads to the zwitterionic germylene BzL′Ge: 5 (BzL′=ArNC(?CH2)C(Bz)?C(Me)NAr) which could be isolated in 83 % yield. The benzyl group has a distinct influence on the reactivity of zwitterionic 5 in comparison to its benzyl‐free analogue, as shown by the reaction of 5 with phenylacetylene, which yields solely the 1,4‐addition product 6, that is, the alkynyl germylene BzLGeCCPh. Compounds 2–8 have been fully characterized by multinuclear NMR spectroscopy, mass spectrometry, elemental analyses, and single‐crystal X‐ray diffraction analyses.  相似文献   

12.
Further investigations into the chemistry of the rhenacyclobutadiene complexes (CO)4Re(η2-C(R)C(CO2Me)C(X)) (1: R=Me, X=OEt (1a), O(CH2)3CCH (1b), NEt2 (1c); R=CHEt2, X=OEt (1d); R=Ph, X=OEt (1e)) are reported. Reactions of 1 with alkynes at reflux temperature of toluene and at ambient temperature either under photochemical conditions or in the presence of PdO yield ring-substituted η5-cyclopentadienylrhenium tricarbonyl complexes, 2. The symmetrical alkynes RCCR (R=Ph, Me, CO2Me) afford the pentasubstituted complexes (η5-C5(Me)(CO2Me)(OEt)(Ph)(Ph))Re(CO)3 (2d), (η5-C5(Me)(CO2Me)(OEt)(Me)(Me))Re(CO)3 (2e), (η5-C5(Me)(CO2Me)(OEt)(CO2Me)(CO2Me))Re(CO)3 (2f), and (η5-C5(Me)(CO2Me)(NEt2)(CO2Me)(CO2Me))Re(CO)3 (2i) on reaction with the appropriate 1, whereas the unsymmetrical alkynes RCCR″ (R=Ph; R″=H, Me) give either only one, (η5-C5(Me)(CO2Me)(OEt)(Ph)H)Re(CO)3 (2a)), or both, (η5-C5(Me)(CO2Me) (OEt)(Ph)(Me))Re(CO)3 (2b) and (η5-C5(Me)(CO2Me)(OEt)(Me)(Ph))Re(CO)3 (2c), (η5-C5(Ph)(CO2Me)(OEt)(Ph)H)Re(CO)3 (2g) and (η5-C5(Ph)(CO2Me)(OEt)(H)(Ph))Re(CO)3 (2h), of the possible products of [3 + 2] cycloaddition of alkyne to η2-C(R)C(CO2Me)C(X). Thermolysis of (CO)4Re(η2-C(Me)C(CO2Me)C(O(CH2)3CCH)) (1b) containing a pendant alkynyl group proceeds to (η5-C5(Me)(CO2Me)(O(CH2)3)H)Re(CO)3 (2j), a η5-cyclopentadienyl-dihydropyran fused-ring product. Competition experiments showed that each of PhCCH and MeO2CCCCO2Me reacts faster than PhCCPh with 1a. The results with unsymmetrical alkynes are rationalized by steric properties of substituents at the CC and ReC bonds and by a preference of ReC(Me) over ReC(OEt) to undergo alkyne insertion. A mechanism is proposed that involves substitution of a trans CO by alkyne in 1, insertion of alkyne into ReC bond to give a rhenabenzene intermediate, and collapse of the latter to 2. Complexes 1a and 1d undergo rearrangement in MeCN at reflux temperature to give rhenafuran-like products, (CO)4Re(κ2-OC(OMe)C(CHCR2)C(OEt)) (R=H (3a) or Et (3b)). The reaction of 1d also proceeds in EtCN, PhCN, and t-BuCN at comparable temperature, but is slower (especially in t-BuCN) than in MeCN. In pyridine at reflux temperature, 1a undergoes a similar rearrangement, with CO substitution, to give (CO)3(py)Re(κ2-OC(OMe)C(CHCEt2)C(OEt)) (4). A mechanism is proposed for these reactions. The sulfonium ylides Me2SCHC(O)Ph and Me2SC(CN)2 (Me2SCRR) react with 1a in acetonitrile at reflux temperature by nucleophilic addition of the ylide to the ReC(Me) carbon, loss of Me2S, and rearrangement to a rhenafuran-type structure to yield (CO)4Re(κ2-OC(OMe)C(C(Me)CRR)C(OEt)) (R=H, R=C(O)Ph (5a); R=RCN (5b)). All new compounds were characterized by a combination of elemental analysis, mass spectrometry, and IR and NMR spectroscopy.  相似文献   

13.
Unusual chemical transformations such as three‐component combination and ring‐opening of N‐heterocycles or formation of a carbon–carbon double bond through multiple C–H activation were observed in the reactions of TpMe2‐supported yttrium alkyl complexes with aromatic N‐heterocycles. The scorpionate‐anchored yttrium dialkyl complex [TpMe2Y(CH2Ph)2(THF)] reacted with 1‐methylimidazole in 1:2 molar ratio to give a rare hexanuclear 24‐membered rare‐earth metallomacrocyclic compound [TpMe2Y(μN,C‐Im)(η2N,C‐Im)]6 ( 1 ; Im=1‐methylimidazolyl) through two kinds of C–H activations at the C2‐ and C5‐positions of the imidazole ring. However, [TpMe2Y(CH2Ph)2(THF)] reacted with two equivalents of 1‐methylbenzimidazole to afford a C–C coupling/ring‐opening/C–C coupling product [TpMe2Y{η3‐(N,N,N)‐N(CH3)C6H4NHCH?C(Ph)CN(CH3)C6H4NH}] ( 2 ). Further investigations indicated that [TpMe2Y(CH2Ph)2(THF)] reacted with benzothiazole in 1:1 or 1:2 molar ratio to produce a C–C coupling/ring‐opening product {(TpMe2)Y[μ‐η21‐SC6H4N(CH?CHPh)](THF)}2 ( 3 ). Moreover, the mixed TpMe2/Cp yttrium monoalkyl complex [(TpMe2)CpYCH2Ph(THF)] reacted with two equivalents of 1‐methylimidazole in THF at room temperature to afford a trinuclear yttrium complex [TpMe2CpY(μ‐N,C‐Im)]3 ( 5 ), whereas when the above reaction was carried out at 55 °C for two days, two structurally characterized metal complexes [TpMe2Y(Im‐TpMe2)] ( 7 ; Im‐TpMe2=1‐methyl‐imidazolyl‐TpMe2) and [Cp3Y(HIm)] ( 8 ; HIm=1‐methylimidazole) were obtained in 26 and 17 % isolated yields, respectively, accompanied by some unidentified materials. The formation of 7 reveals an uncommon example of construction of a C?C bond through multiple C–H activations.  相似文献   

14.
The reactivity of a series of iridium? pyridylidene complexes with the formula [TpMe2Ir(C6H5)2(C(CH)3C(R)N H] ( 1 a – 1 c ) towards a variety of substrates, from small molecules, such as H2, O2, carbon oxides, and formaldehyde, to alkenes and alkynes, is described. Most of the observed reactivity is best explained by invoking 16 e? unsaturated [TpMe2Ir(phenyl)(pyridyl)] intermediates, which behave as internal frustrated Lewis pairs (FLPs). H2 is heterolytically split to give hydride? pyridylidene complexes, whilst CO, CO2, and H2C?O provide carbonyl, carbonate, and alkoxide species, respectively. Ethylene and propene form five‐membered metallacycles with an IrCH2CH(R)N (R=H, Me) motif, whereas, in contrast, acetylene affords four‐membered iridacycles with the IrC(?CH2)N moiety. C6H5(C?O)H and C6H5C?CH react with formation of Ir? C6H5 and Ir? C?CPh bonds and the concomitant elimination of a molecule of pyridine and benzene, respectively. Finally the reactivity of compounds 1 a – 1 c against O2 is described. Density functional theory calculations that provide theoretical support for these experimental observations are also reported.  相似文献   

15.
A novel, useful in situ synthesis for NHC nickel allyl halide complexes [Ni(NHC)(η3-allyl)(X)] starting from [Ni(CO)4], NHC and allyl halides is presented. The reaction of [Ni(CO)4] with (i) one equivalent of the corresponding NHC and (ii) with an excess of the corresponding allyl chloride at room temperature leads with elimination of carbon monoxide to complexes of the type [Ni(NHC)(η3-allyl)(X)]. This approach was used to synthesize the complexes [Ni(tBu2Im)(η3-H2C -C (Me)-C H2)(Cl)] ( 2 ), [Ni(iPr2ImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 3 ), [Ni(iPr2Im)(η3-H2C -C (Me)-C H2)(Cl)] ( 4 ), [Ni(iPr2Im)(η3-H2C -C (H)-C (Me)2)(Br)] ( 5 ), [Ni(Me2ImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 6 ), and [Ni(EtiPrImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 7 ). The complexes 1 to 7 were characterized using NMR and IR spectroscopy and elemental analysis, and the molecular structures are provided for 2 and 7 . The allyl nickel complexes 1 – 7 are stereochemically non-rigid in solution due to (i) NHC rotation about the nickel-carbon bond, (ii) allyl rotation about the Ni–η3-allyl axis and (iii) π–σ–π allyl isomerization processes. The allyl halide complexes can be methylated as was demonstrated by the methylation of a number of the complexes [Ni(NHC)(η3-allyl)(X)] with methylmagnesium chloride or methyllithium, which led to isolation of the complexes [Ni(Me2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 8 ), [Ni(tBu2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 9 ), [Ni(iPr2ImMe)(η3-H2C -C (Me)-C H2)(Me)] ( 10 ), [Ni(iPr2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 11 ), [Ni(iPr2Im)(η3-H2C -C (H)-C (Me)2)(Me)] ( 12 ), and [Ni(EtiPrImMe)(η3-H2C -C (Me)-C H2)(Me)] ( 13 ). These complexes were fully characterized including X-ray molecular structures for 10 and 11 .  相似文献   

16.
The rhenacyclobutadienes (CO)4Re(η2- C(R)C(CO2Me)C(OR)) (2) undergo a number of reactions that mirror those of Fischer alkoxycarbene complexes. Thus, (CO)4Re(η2-C(Me)C(CO2Me)C(OEt)) (2a) can be deprotonated by LDA, Na[OBu-t], or Na[CH(CO2Me)2] to give the ylide-like conjugate base [(CO)4Re(η2-C(CH2)C(CO2Me)C(OEt)] (3), which was isolated as PPN(3). Li(3) undergoes deuteriation with DCl/D2O and alkylation with Et3OPF6 at ReCCH2, with the latter reaction affording (CO)4Re(η2-C(CH2Et)C(CO2Me)C(OEt)) (4). Repetition of the sequence deprotonation-ethylation on 4 generates (CO)4Re(η2-C(CHEt2)C(CO2Me)C(OEt)) (5). The nature of the alkoxy substituent in 2 can be varied by use of the rhenacyclobutenones Na[(CO)4Re(η2-C(R)C(CO2Me)C(O))] (Na(1)) in conjunction with AcCl and ROH to produce a series of new complexes (R=Ph, R=Et (2b); R=Me, R=CH2CHCH2 (2c), (CH2)3CCH (2d), Me (2e)). Aminolysis of 2a with the primary and secondary amines PhNH2, HO(CH2)2NH, p-TolNH2, and Et2NH yields the aminorhenacyclobutadiene complexes (CO)4Re(η2-C(Me)C(CO2Me)C(NHR or NR2)) (R2=Et2 (6a); R=Ph (6b), (CH2)2OH (6c), p-Tol (6d)). These complexes display lesser carbene-like character than their alkoxy counterparts 2, as evidenced by 1H and 13C NMR spectroscopic properties and lack of reactivity toward LDA by 6a. Reactions of each 2a and 6a with PPhMe2 at low temperature afford (CO)4Re(η2-C(Me)(PPhMe2)C(CO2Me)C(OEt)) (7) and (CO)3(PPhMe2)Re(η2-C(Me)C(CO2Me)C(NEt2)) (9), respectively, further in agreement with the more carbenoid nature of 2a than 6a. 7 undergoes conversion to (CO)3(PPhMe2)Re(η2-C(Me)C(CO2Me)C(OEt)) (8) upon heating. 2a reacts with each of (NH4)2[Ce(NO3)6], DMSO, EtNO2/Et3N, and Me3NO under various conditions to afford one or both of the oxygen atom insertion products into the ReC bonds, (CO)4Re(κ2-OC(Me)C(CO2Me)C(OEt)) (10) and (CO)4Re(κ2-C(Me)C(CO2Me)C(OEt)O) (11). In contrast, no reaction occurred between 2a and S8 on heating. However, 6a was converted to the NH insertion product (CO)4Re(κ2-NHC(Me)C(CO2Me)C(NEt2)) (12) by the action of H2NNH2 · H2O at 0 °C. All new compounds were characterized by a combination of elemental analysis, mass spectrometry, and IR and NMR spectroscopy.  相似文献   

17.
Deprotonation of methyl acetoacetate yields two enolate ions MeCOC?HCO2Me (a) and C?H2COCH2CO2Me (b). On collisional activation, ions a and b fragment differently. The major fragmentation of a is specific loss of MeOH through a four-centred transition state to form ?O(Me)C?C?C?O. In contrast, ion b eliminates CH2CO to give ?CH2CO2Me. Some rearrangement of b to a is also noted. Rearrangement of a to b is very minor under single collision conditions but at high collision gas pressure rearrangement of a to b is strongly promoted. Similar effects are observed in the collisional activation spectra of MeCOC?(Me)CO2Me (c) and ?CH2COC(Me)CO2Me (d). The loss of MeOH from (c) proceeds via a six membered transition state to ?CH2? CO? C(Me)?C?O; this is a stepwise process in which the deprotonation (step two) is not rate determining. A number of other decompositions occur, these have also been studied by deuterium labelling.  相似文献   

18.
Iridabicycles [Ir{κ3-N,C,O-(pyC(H)=C(C(O)Me)2}(Cl)(L−L)](L−L=cod (cod=1,5-cyclooctadiene), 1 a ; bipy (bipy=2,2’-bipyridine), 1 b ) have been obtained by oxidative coordination of 3-(pyridine-2-yl-methylene)pentane-2,4-dione L1 , to the complexes [{Ir(μ-Cl)(cod)}2] and [{Ir(μ-Cl)(coe)2}2] (coe=cis-cyclooctene), the latter in the presence of bipy. Remarkably, cleavage of the C3−C(O)Me bond of L1 has instead been achieved in the reaction with [Ir(Cl)(dmb)2] (dmb=2,3-dimethylbutadiene), yielding a compound formulated as [Ir{κ2-N,C-(pyC(H)C(C(O)Me))}(CO)(μ-Cl)(Me)]2, 2 . Treatment of dimer 2 with DMSO or PMe3 produced the complexes[Ir{κ2-N,C-(pyC(H)C(C(O)Me)}(CO)Cl(Me)L] (L=DMSO, 3 a ; PMe3, 3 b ). Plausible mechanisms for the reactions leading to complexes 1 and 2 are proposed by means of DFT calculations.  相似文献   

19.
The interaction of Ph3Ti with a number of ketones (acetone, 2-butanone, 3,3-dimethyl-2-butanone, benzophenone) was studied. The PhTi(OCR1R2Ph)2 complexes, where R1=R2=Me (a), R1=Me, R2=Et (b), R1=Me, R2=t-Bu (c), and R1=R2=Ph (d) were isolated in satisfactory yields. These compounds were characterized by ESR and IR spectroscopy and elemental analysis. Their thermal stability was determined by the DTA method. The reaction of Bn3V with acetone gives V(OCMe2Bn)3. In analogy with titanium compounds, Bn4V reacts with acetone at the ratio 12 to give Bn2(OCMe2Bn)2.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1120–1122, June, 1994.  相似文献   

20.
A series of unusual chemical‐bond transformations were observed in the reactions of high active yttrium? dialkyl complexes with unsaturated small molecules. The reaction of scorpionate‐anchored yttrium? dibenzyl complex [TpMe2Y(CH2Ph)2(thf)] ( 1 , TpMe2=tri(3,5‐dimethylpyrazolyl)borate) with phenyl isothiocyanate led to C?S bond cleavage to give a cubane‐type yttrium–sulfur cluster, {TpMe2Y(μ3‐S)}4 ( 2 ), accompanied by the elimination of PhN?C(CH2Ph)2. However, compound 1 reacted with phenyl isocyanate to afford a C(sp3)? H activation product, [TpMe2Y(thf){μ‐η13‐OC(CHPh)NPh}{μ‐η32‐OC(CHPh)NPh}YTpMe2] ( 3 ). Moreover, compound 1 reacted with phenylacetonitrile at room temperature to produce γ‐deprotonation product [(TpMe2)2Y]+[TpMe2Y(N=C?CHPh)3]? ( 6 ), in which the newly formed N?C?CHPh ligands bound to the metal through the terminal nitrogen atoms. When this reaction was carried out in toluene at 120 °C, it gave a tandem γ‐deprotonation/insertion/partial‐TpMe2‐degradation product, [(TpMe2Y)2(μ‐Pz)2{μ‐η13‐NC(CH2Ph)CHPh}] ( 7 , Pz=3,5‐dimethylpyrazolyl).  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