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 468 毫秒
1.
The reactions of methyl and ethyl esters of valine with m-bromobenzaldehyde and p-chlorobenzaldehyde in absolute methanol in the presence of magnesium sulfate yielded the corresponding azomethines (CH3)2CHCH(COOR1)N=CHR2 (R1 = CH3, C2H5; R2 = 3-BrC6H4, 4-ClC6H4); their reduction with sodium borohydride gave N-benzyl derivatives of valine esters, (CH3)2CHCH(COOR1)NHCH2R2.Translated from Zhurnal Obshchei Khimii, Vol. 74, No. 11, 2004, pp. 1855–1857.Original Russian Text Copyright © 2004 by Yakubovich, Zhavnerko, Shirokii, Knizhnikov.This revised version was published online in April 2005 with a corrected cover date.  相似文献   

2.
Preparation of Enantiomerically Pure Derivatives of 3-Amino- and 3-Mercaptobutanoic Acid by SN2 Ring Opening of the β-Lactone and a 1,3-Dioxanone Derived from 3-Hydroxybutanoic Acid From (S)-4-methyloxetan-2-one ( 1 ), the β-butyrolactone readily available from the biopolymer ( R )-polyhydroxybutyrate (PHB) and various C, N, O and S nucleophiles, the following compounds are prepared:(s)-2-hydroxy-4-octanone ( 3 ), (R)-3-aminobutanoic acid ( 7 ) and its N-benzyl derivative 5 , (R)-3-azidobutanoic acid ( 6 ) (R)-3-mercaptobutanoic acid ( 10 ), (R)-3-(phenylthio)butanoic acid ( 8 ) and its sulfoxide 9 . The (6R)-2,6-dimethyl-2-ethoxy-1,3-dioxan-4-one ( 4 ) from (R)-3-hydroxybutanoic acid undergoes SN2 ring opening with benzylamine to give the N-benzyl derivative (ent- 5 ) of (S)-3-aminobutanoic acid in 30?40% yield.  相似文献   

3.
The magnitude of chemical shift nonequivalence (Δδ) in compounds having the general formula R1CH(COOR2)N+R3Me2X? is discussed from the viewpoint of selective shielding of one of the geminal groups by the ester group in the preferred rotamer. In addition, a novel example of the increase of Δδ with rise in temperature has been observed.  相似文献   

4.
Treatment of N,N‐chelated germylene [(iPr)2NB(N‐2,6‐Me2C6H3)2]Ge ( 1 ) with ferrocenyl alkynes containing carbonyl functionalities, FcC≡CC(O)R, resulted in [2+2+2] cyclization and formation of the respective ferrocenylated 3‐Fc‐4‐C(O)R‐1,2‐digermacyclobut‐3‐enes 2 – 4 [R = Me ( 2 ), OEt ( 3 ) and NMe2 ( 4 )] bearing intact carbonyl substituents. In contrast, the reaction between 1 and PhC(O)C≡CC(O)Ph led to activation of both C≡C and C=O bonds producing bicyclic compound containing two five‐membered 1‐germa‐2‐oxacyclopent‐3‐ene rings sharing one C–C bond, 4,8‐diphenyl‐3,7‐dioxa‐2,6‐digermabicyclo[3.3.0]octa‐4,8‐diene ( 5 ). With N‐methylmaleimide containing an analogous C(O)CH=CHC(O) fragment, germylene 1 reacted under [2+2+2] cyclization involving the C=C double bond, producing 1,2‐digermacyclobutane 6 with unchanged carbonyl moieties. Finally, 1 selectively added to the terminal double bond in allenes CH2=C=CRR′ giving rise to 3‐(=CRR′)‐1,2‐digermacyclobutanes [R/R′ = Me/Me ( 7 ), H/OMe ( 8 )] bearing an exo‐C=C double bond. All compounds were characterized by 1H, 13C{1H} NMR, IR and Raman spectroscopy and the molecular structures of 3 , 4 , 5 , and 8 were established by single‐crystal X‐ray diffraction analysis. The redox behavior of ferrocenylated derivatives 2 – 4 was studied by cyclic voltammetry.  相似文献   

5.
Abstract

Reaction of two equivalents of N-mono- or di-substituted 3-amino-4-(n-butoxy)-3-cyclobutene-1,2-diones with a 1,2-diaminoethane gave N-mono- or di-substituted 1,2-bis((2-amino-1-cyclobutene-3,4-dione)amino)-ethane derivatives (bis(squaramides)). Reaction of the bis(squaramides) with excess P4S10 gave the analogous tetrathio derivatives (bis(dithiosquaramides), LH2) of formula (NR1R2)C4S2(NHCH2CH2NH)-C4S2(NR1R2) (R1=n-Bu, R2=H; R1=R2=Et, n-Bu). The new bis(dithiosquaramide) ligands were characterized by elemental analysis, IR, 1H NMR, 13C NMR, electronic, and mass spectroscopic methods. The complexes of these ligands with nickel(II) were prepared, isolated and characterized. The isolated complexes are neutral 2:2 species of formula Ni2L2, as evidenced by results from mass spectrometry, and they exhibit thermochromic behaviour in pyridine solution. Additional spectroscopic data (IR, NMR) are consistent with the ligands being coordinated only through sulfur donor atoms and a structure for the complexes is proposed.  相似文献   

6.
Optically active mixed alkoxy orthotitanates with general formula Ti(OR1)2(OR2)(OR3) (R1=Et, Bun; R2=CH2CH2OCOC(Me)=CH2; R3=menthyl, CH(Me)CH2Me, CH(Ph)CH(NHMe)Me, CH(C9H6N)(C9H14N)) were obtained for the first time by transesterification. The TiIV monomers synthesized were characterized by elemental analysis, ozonolysis, and1H and13C NMR and IR spectroscopy. Polymer products with optical activity were obtained by liquid phase radical copolymerization of TiIV-containing monomers. For Part 51, see Ref. 1. Deceased. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1739–1743, September, 1999.  相似文献   

7.
The allyl Mn(II) organic compounds R1CH=C(R2)CH2MnCl (R1=H, Me; R2=H, Me, Bu), obtained in situ from Grignard reagents and Li2MnCl4, react with esters of 4-bromocrotonic, (2-bromobutylidene)-, (4-bromo-2-butenylidene)-, (2-bromoisobutylidene) malonic, and (2-bromoheptylidene)cyanoacetic acids in THF at –78 to +20C to give derivatives of substituted cyclopropanecarboxylic or cyclopropane-1, 1-dicarboxylic acids. These derivatives contain a fragment of the allyl type. When ethers of 2-(bromomethyl)acrylic, 4-bromo-2-methyl-, and 4-bromo-3-methyl-2-butenoic acids are used, cross-combination products result.For previous communication, see [1].Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 3, pp. 656–663, March, 1991.  相似文献   

8.
Summary The mechanism of the oxidation of mixtures of 2,6-dimethylaniline (1), N,N-dimethylaniline (2), 2,6-diethylaniline (3), N,N-diethylaniline (4), N-methylaniline (5), 2,6-difluoroaniline (6), and 2,3,5,6-tetrafluoroaniline (7) with 4-aminophenol (8) by cerium(IV) ions in aqueous perchloric acid has been investigated. The indoaniline salts [O=C6H4=N-C6H2(R 1)2NH(R 2)2]+ClO 4 (R 1=H,R 2=CH3, C2H5 orvice versa) are formed as intermediates in the cross-coupling reaction; they undergo oxidation to imino-4-benzoquinone (9) and its corresponding derivatives by cerium(IV) ions in high yields. The mechanism of this process is discussed.
Durch Cer(IV)-Ionen induzierte oxidative Kreuzkupplung einiger 2,6- und N,N-disubstituierter Anilinderivate mit 4-Aminophenol in wässriger Perchlorsäure
Zusammenfassung Die Oxidation von Mischungen von 2,6-Dimethylanilin (1), N,N-Dimethylanilin (2), 2,6-Diethylanilin (3), N,N-Diethylanilin (4), N-Methylanilin (5), 2,6-Difluoranolin (6) und 2,3,5,6-Tetrafluoranilin (7) mit 4-Aminophenol (8) durch Cer(IV)-Ionen in wässriger Perchlorsäure wurde untersucht. Als Zwischenprodukte der Kreuzkupplungsreaktion treten die Indoanilinsalze [O=C6H4=N-C6H2(R 1)2NH(R 2)2]+ClO 4 (R 1=H,R 2=CH3, C2H5 oder umgekehrt) auf. Diese werden durch Cer(IV)-Ionen in hohen Ausbeuten zu Imino-4-benzochinon (9) und seinen entsprechenden Derivaten oxidiert. Der Mechanismus dieses Vorgangs wird diskutiert.
  相似文献   

9.
Abstract

The reactions of a variety of electrophiles with the N-silyl-P-trifluoroethoxyphosphoranimine anion Me3Sin°P(Me)(OCH2CF3)CH? 2 (1a), prepared by the deprotonation of the dimethyl precursor Me3SiN[dbnd]P(OCH2CF3)Me2 (1) with n-BuLi in Et2O at-78°C, were studied. Thus, treatment of 1a with alkyl halides, ethyl chloroformate, or bromine afforded the new N-silylphosphoranimine derivatives Me3SiN[dbnd]P(Me)(OCH2CF3)CH2R [2: R = Me, 3: R = CH2Ph, 4: R = CH[sbnd]CH2, 5: R = C(O)OEt, and 6: R = Br]. In another series, when 1a was allowed to react with various carbonyl compounds, 1,2-addition of the anion to the carbonyl group was observed. Quenching with Me3SiCl gave the O-silylated products Me3SiN[dbnd]P(Me)(OCH2CF3)CH2°C(OSiMe3)R1R2 [7: R 1 = R 2 = Me; 8: R 1 = Me, R 2 = Ph; 9: R1 = Me, R 2 = CH[sbnd]CH2; and 10: R 1 = H, R 2 = Ph]. Compounds 2–10 were obtained as distillable, thermally stable liquids and were characterized by NMR spectroscopy (1H, 13C, and 31P) and elemental analysis.  相似文献   

10.
Sequential treatment of 2‐C6H4Br(CHO) with LiC≡CR1 (R1=SiMe3, tBu), nBuLi, CuBr?SMe2 and HC≡CCHClR2 [R2=Ph, 4‐CF3Ph, 3‐CNPh, 4‐(MeO2C)Ph] at ?50 °C leads to formation of an intermediate carbanion (Z)‐1,2‐C6H4{CA(=O)C≡CBR1}{CH=CH(CH?)R2} ( 4 ). Low temperatures (?50 °C) favour attack at CB leading to kinetic formation of 6,8‐bicycles containing non‐classical C‐carbanion enolates ( 5 ). Higher temperatures (?10 °C to ambient) and electron‐deficient R2 favour retro σ‐bond C?C cleavage regenerating 4 , which subsequently closes on CA providing 6,6‐bicyclic alkoxides ( 6 ). Computational modelling (CBS‐QB3) indicated that both pathways are viable and of similar energies. Reaction of 6 with H+ gave 1,2‐dihydronaphthalen‐1‐ols, or under dehydrating conditions, 2‐aryl‐1‐alkynylnaphthlenes. Enolates 5 react in situ with: H2O, D2O, I2, allylbromide, S2Me2, CO2 and lead to the expected C ‐E derivatives (E=H, D, I, allyl, SMe, CO2H) in 49–64 % yield directly from intermediate 5 . The parents (E=H; R1=SiMe3, tBu; R2=Ph) are versatile starting materials for NaBH4 and Grignard C=O additions, desilylation (when R1=SiMe) and oxime formation. The latter allows formation of 6,9‐bicyclics via Beckmann rearrangement. The 6,8‐ring iodides are suitable Suzuki precursors for Pd‐catalysed C?C coupling (81–87 %), whereas the carboxylic acids readily form amides under T3P® conditions (71–95 %).  相似文献   

11.
Synthesis and Spectroscopic Characterisation of some Pentacarbonyltungsten(0) Complexes with Mono‐ and Bicyclic Phosphirane Ligands: Crystal Structure of [{(Me3Si)2HCPC(H)H–C(H)Ph}W(CO)5] The tungsten(0) complex [{(Me3Si)2HCPC(Ph)=N}W(CO)5] ( 1 ) reacts upon heating with alkene derivatives 2 , 6 , 8 , and 10 in toluene to form benzonitrile and the complexes [{(Me3Si)2HCPC(R1,R2)–C(R3,R4}W(CO)5] ( 4 , 7 a , b , 9 a , b , 11 a , b ) ( 4 (trans): R1,R3 = Ph, R2,R4 = H, 7 a , b (cis, meso and rac): R1,R3 = Ph, R2,R4 = H, 9 a , b (RR und SS): R1 = Ph, R2,R3,R4 = H, 11 a , b : R1=R3 = (CH2)4, R2,R4 = H). Spectroscopic and mass spectrometric data are discussed. The structure of the complex 9 a was determined by X‐ray single crystal structure analysis showing characteristic data for the phosphirane ring such as a narrow angle at phosphorus (49,2(2)°), different P–C distances (P–C(6) 182,1(5) and P–C(7) 185,2(4) pm) and 152,9(6) pm for the basal C–C bond.  相似文献   

12.
The seven rhenium (I) tricarbonyl complexes having a general formula fac‐[ReBr(CO)3(R1,R2,R3‐N^N)] (N^N = imidazo[4,5‐f]‐1,10‐phenanthroline; R1 = ? tBu, R2 = R3 = ? H, 1 ; R1 = ? C?CH, R2 = R3 = ? H, 2 ; R1 = ? tBu, R2 = ? C?CH, R3 = ? H, 3 ; R1 = ? tBu, R2 = R3 = ? C?CH, 4 ; R1 = ? tBu, R2 = ? CH3, R3 = ? H, 5 ; R1 = ? tBu, R2 = R3 = ? CH3, 6 ; R1 = ? tBu, R2 = ? OCH3, R3 = ? H, 7 ) have been investigated theoretically by density functional theory (DFT) and time‐dependent density functional theory (TDDFT) methods. The different substituted groups on N^N ligand induce changes on the electronic structures and photophysical properties for these complexes. It is found that the introduction of ? C?C decreases the energy level of lowest unoccupied molecular orbital (LUMO) while the introduction of ? CH3 or ? OCH3 lead to increase the energy level of LUMO. The order of LUMO energy level rising is in line with the increasing of donating abilities of substituted groups; and the influence of R2 position is greater than that of R1 position on LUMO energy level. The lowest energy absorption bands have changes in the order of 7 < 6 < 5 < 1 < 2 < 3 < 4 . These results of electronic affinity (EA), ionization potential (IP), and reorganization energy (λ) indicate that all of these complexes can be used as electron transporting materials. Moreover, the smallest difference between λelectron and λhole of 4 indicates that it is better to be used as an emitter in the organic light‐emitting diodes. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
The aprotic and protic bi- and multidentate iminophosphines 2-Ph2PC6H4N=CR1R2 (R1=H, R2=Ph=2a; R1=Me R2=Ph=2b; R1=H, R2=2-thienyl=2c; R1=H, R2=C6H4-2-PPh2=2d; R1=H, R2=C6H4-2-OH=2e, R1=H, R2=C6H4-2-OH-3-But=2f; R1=H, R2=CH2C(O)Me=2g) have been prepared by the acid catalyzed condensation of 2-(diphenylphosphino)aniline with the corresponding aldehyde–ketone. Iminophosphine 2d can be reduced with sodium cyanoborohydride to give the corresponding amino-diphosphine 2-Ph2PC6H4N(H)CH2C6H4-2-PPh2 (2h). In the presence of a stoichiometric quantity of acid, 2-(diphenylphosphino)aniline reacts in an unexpected manner with benzaldehyde, salicylaldehyde, or acetophenone to give the corresponding 2,3-dihydro-1H-benzo[1,3]azaphosphol-3-ium salts and with pyridine-2-carboxaldehyde to give N-(pyridin-2-ylmethyl)-2-diphenylphosphinoylaniline, the latter of which has been characterized by single-crystal X-ray crystallography, as its palladium dichloride derivative. The attempted condensation of 2-(diphenylphosphino)aniline with pyridine-2-carboxaldehyde to give the corresponding pyridine-functionalized iminophosphine resulted in an unusual transformation involving the diastereoselective addition of two equivalents of aldehyde to give 1,2-dipyridin-2-yl-2-(o-diphenylphosphinoyl)phenylamino-ethanol, which has been characterized by a single-crystal X-ray structure determination. The bidentate iminophosphine 2-Ph2PC6H4N=C(H)Ph reacts with [(cycloocta-1,5-diene)PdClX] X=Cl, Me) to give [Pd{2-Ph2PC6H4N=C(H)Ph}ClX] and the imino-diphosphine 2-Ph2PC6H4N=C(H)C6H4-PPh2 reacts with [(cycloocta-1,5-diene)PdClMe] to give [Pd{2-Ph2PC6H4N=C(H)C6H4---PPh2}ClMe] and each has been characterized by single-crystal X-ray crystallography. The monobasic iminophosphine 2-Ph2PC6H4N=C(Me)CH2C(O)Me reacts with [Ni(PPh3)2Cl2] in the presence of NaH to give the phosphino–ketoiminate complex [Ni{2-Ph2PC6H4N=C(Me)CHC(O)Me}Cl], which has been structurally characterized. Mixtures of iminophosphines 2ah and a palladium source catalyze the Suzuki cross coupling of 4-bromoacetophenone with phenyl boronic acid. The efficiency of these catalysts show a marked dependence on the palladium source, catalysts formed from [Pd2(OAc)6] giving consistently higher conversions than those formed from [Pd2(dba)3] and [PdCl2(MeCN)2]. Catalysts formed from neutral bi- and terdentate iminophosphines 2ad gave significantly higher conversions than those formed from their monobasic counterparts 2ef. Notably, under our conditions the conversions obtained with 2ac compare favorably with those of the standards; catalysts formed from tris(2-tolyl)phosphine and tris(2,4-di-tert-butylphenyl)phosphite and a source of palladium. In addition, mixtures of [Ir(COD)Cl]2 and 2ah are active for the hydrosilylation of acetophenone; in this case catalysts formed from monobasic iminophosphines 2ef giving the highest conversions.  相似文献   

14.
《Tetrahedron: Asymmetry》1998,9(23):4219-4238
A wide variety of planar chiral cyclopalladated compounds of general formulae [Pd{[(η5-C5H3)–CHN–CH(Me)–C10H7]Fe(η5-C5H5)}Cl(L)] (with L=py-d5 or PPh3), [Pd{[(η5-C5H3)–CHN–CH(Me)–C10H7]Fe(η5-C5H5)}(acac)] or [Pd{[(R1–CC–R2)25-C5H3)–CHN–CH(Me)–C10H7]Fe(η5-C5H5)}Cl] (with R1=R2=Et; R1=Me, R2=Ph; R1=H, R2=Ph; R1=R2=Ph; R1=R2=CO2Me or R1=CO2Et, R2=Ph) are reported. The diastereomers {(Rp,R) and (Sp,R)} of these compounds have been isolated by either column chromatography or fractional crystallization. The free ligand (R)-(+)-[{(η5-C5H4)–CHN–CH(Me)–C10H7}Fe(η5–C5H5)] (1) and compound (+)-(Rp,R)-[Pd{[(Et–CC–Et)25-C5H3)–CHN–CH(Me)–C10H7]Fe(η5-C5H5)}Cl] (7a) have also been characterized by X-ray diffraction. Electrochemical studies based on cyclic voltammetries of all the compounds are also reported.  相似文献   

15.
(Diphenylphosphanyl)phenols C6H3(1‐OH)(2‐PPh2)(4‐R1)(6‐R2), abbreviated as (POH), oxidatively add to Fe(PMe3)4 affording hydridoiron(II) compounds fac‐FeH(PO)(PMe3)3 ( 1 : R1=R2=H; 2 : R1=Me, R2=H; 3 : R1=OMe, R2=H; 4 : R1=Me, R2=CMe3; 5 : R1=R2=CMe3) with high stereoselectivity. (2‐diphenylphosphanyl)thiophenol (PSH) reacts accordingly forming fac‐FeH(PS)(PMe3)3 ( 9 ). Complete assignment of 1H, 13C, and 31P signals is achieved by 2D heteronuclear shift correlations. 4,6‐Di‐tert‐butyl‐(2‐diphenylphosphanyl)phenol reacts with FeI(Me)(PMe3)4 to form FeI(PO)(PMe3)2 ( 6 ). 4 , 5 and 9 under 1 bar of CO are converted to monocarbonyl derivatives FeH(PX)(CO)(PMe3)2 ( 7 , 8 : X = O; 10 : X = S) which in solution form mixtures of two isomers A and B . 4 and 5 react with their parent phosphanylphenols, respectively, to give diamagnetic complexes Fe(PO)2(PMe3) ( 11 , 12 ) which dissociate trimethylphosphane to give paramagnetic compounds Fe(PO)2. The same phosphanylphenols react with FeCl3 to afford racemic mixtures of complexes Fe(PO)3 ( 13 , 14 ). Structural data were also obtained from single crystals of compounds 1 , 5 , and 11 .  相似文献   

16.
Two series of 5-trichloromethylisoxazoles were synthesized from the cyclocondensation of 1,1,1-trichloro-4-methoxy-3-alken-2-ones [Cl3CC(O)C(R2) = C(R1)OMe, where R1 = H, Me, Et, Pr, iso-Pr, cyclo-Pr, Bu, terc-Bu, CH2Br, CHBr2, CH(Me)SMe, (CH2)2Ph, and Ph, and R2 = H; R1 = H and R2 = Me and Et; R1 and R2 = -(CH2)4- and -(CH2)5-; and R1 = Et and Ph and R2 = Me] with hydroxylamine hydrochloride through a rapid one-pot reaction in water. The 5-trichloromethyl-4,5-dihydroisoxazoles were aromatized by reaction with concentrated sulfuric acid to obtain the respective 5-trichloromethylisoxazoles. Their structures were confirmed by elemental analysis, 1H/13C nuclear magnetic resonance, and electron impact mass spectroscopy. Crystal structure analysis for 5-triclhoromethyl-5-hydroxy-3-propyl-4,5-dihydroisoxazole (2d) and 5-trichloromethyl-5-hydroxy-3,4-hexamethylene-4,5-dihydroisoxazole (2o) is presented. The antimicrobial activities of the 5-trichloromethyl-4,5-dihydroisoxazole derivatives were examined using the standard twofold dilution method against Gram-positive bacteria (Staphylococcus aureus), Gram-negative bacteria (Escherichia coli and Pseudomonas aeruginosa), and yeasts (Candida spp. and Cryptococcus neoformans). All of the tested 5-trichloromethyldihydroisoxazoles exhibited antibacterial and antifungal activities at the tested concentrations.

Supplemental materials are available for this article. Go to the publisher's online edition of Synthetic Communications® to view the free supplemental file.  相似文献   

17.
Treatment of the thioether‐substituted secondary phosphanes R2PH(C6H4‐2‐SR1) [R2=(Me3Si)2CH, R1=Me ( 1PH ), iPr ( 2PH ), Ph ( 3PH ); R2=tBu, R1=Me ( 4PH ); R2=Ph, R1=Me ( 5PH )] with nBuLi yields the corresponding lithium phosphanides, which were isolated as their THF ( 1 – 5Pa ) and tmeda ( 1 – 5Pb ) adducts. Solid‐state structures were obtained for the adducts [R2P(C6H4‐2‐SR1)]Li(L)n [R2=(Me3Si)2CH, R1=nPr, (L)n=tmeda ( 2Pb ); R2=(Me3Si)2CH, R1=Ph, (L)n=tmeda ( 3Pb ); R2=Ph, R1=Me, (L)n=(THF)1.33 ( 5Pa ); R2=Ph, R1=Me, (L)n=([12]crown‐4)2 ( 5Pc )]. Treatment of 1PH with either PhCH2Na or PhCH2K yields the heavier alkali metal complexes [{(Me3Si)2CH}P(C6H4‐2‐SMe)]M(THF)n [M=Na ( 1Pd ), K ( 1Pe )]. With the exception of 2Pa and 2Pb , photolysis of these complexes with white light proceeds rapidly to give the thiolate species [R2P(R1)(C6H4‐2‐S)]M(L)n [M=Li, L=THF ( 1Sa , 3Sa – 5Sa ); M=Li, L=tmeda ( 1Sb , 3Sb – 5Sb ); M=Na, L=THF ( 1Sd ); M=K, L=THF ( 1Se )] as the sole products. The compounds 3Sa and 4Sa may be desolvated to give the cyclic oligomers [[{(Me3Si)2CH}P(Ph)(C6H4‐2‐S)]Li]6 (( 3S )6) and [[tBuP(Me)(C6H4‐2‐S)]Li]8 (( 4S )8), respectively. A mechanistic study reveals that the phosphanide–thiolate rearrangement proceeds by intramolecular nucleophilic attack of the phosphanide center at the carbon atom of the substituent at sulfur. For 2Pa / 2Pb , competing intramolecular β‐deprotonation of the n‐propyl substituent results in the elimination of propene and the formation of the phosphanide–thiolate dianion [{(Me3Si)2CH}P(C6H4‐2‐S)]2?.  相似文献   

18.
The reactions of MCl5 or MOCl3 with imidazole‐based pro‐ligand L1H, 3,5‐tBu2‐2‐OH‐C6H2‐(4,5‐Ph21H‐)imidazole, or oxazole‐based ligand L2H, 3,5‐tBu2‐2‐OH‐C6H2(1H‐phenanthro[9,10‐d])oxazole, following work‐up, afforded octahedral complexes [MX(L1, 2)], where MX=NbCl4 (L1, 1 a ; L2, 2 a ), [NbOCl2(NCMe)] (L1, 1 b ; L2, 2 b ), TaCl4 (L1, 1 c ; L2, 2 c ), or [TaOCl2(NCMe)] (L1, 1 d ). The treatment of α‐diimine ligand L3, (2,6‐iPr2C6H3N?CH)2, with [MCl4(thf)2] (M=Nb, Ta) afforded [MCl4(L3)] (M=Nb, 3 a ; Ta, 3 b ). The reaction of [MCl3(dme)] (dme=1,2‐dimethoxyethane; M=Nb, Ta) with bis(imino)pyridine ligand L4, 2,6‐[2,6‐iPr2C6H3N?(Me)C]2C5H3N, afforded known complexes of the type [MCl3(L4)] (M=Nb, 4 a ; Ta, 4 b ), whereas the reaction of 2‐acetyl‐6‐iminopyridine ligand L5, 2‐[2,6‐iPr2C6H3N?(Me)C]‐6‐Ac‐C5H3N, with the niobium precursor afforded the coupled product [({2‐Ac‐6‐(2,6‐iPr2C6H3N?(Me)C)C5H3N}NbOCl2)2] ( 5 ). The reaction of MCl5 with Schiff‐base pro‐ligands L6H–L10H, 3,5‐(R1)2‐2‐OH‐C6H2CH?N(2‐OR2‐C6H4), (L6H: R1=tBu, R2=Ph; L7H: R1=tBu, R2=Me; L8H: R1=Cl, R2=Ph; L9H: R1=Cl, R2=Me; L10H: R1=Cl, R2=CF3) afforded [MCl4(L6–10)] complexes (M=Nb, 6 a – 10 a ; M=Ta, 6 b – 9 b ). In the case of compound 8 b , the corresponding zwitterion was also synthesised, namely [Ta?Cl5(L8H)+] ? MeCN ( 8 c ). Unexpectedly, the reaction of L7H with TaCl5 at reflux in toluene led to the removal of the methyl group and the formation of trichloride 7 c [TaCl3(L7‐Me)]; conducting the reaction at room temperature led to the formation of the expected methoxy compound ( 7 b ). Upon activation with methylaluminoxane (MAO), these complexes displayed poor activities for the homogeneous polymerisation of ethylene. However, the use of chloroalkylaluminium reagents, such as dimethylaluminium chloride (DMAC) and methylaluminium dichloride (MADC), as co‐catalysts in the presence of the reactivator ethyl trichloroacetate (ETA) generated thermally stable catalysts with, in the case of niobium, catalytic activities that were two orders of magnitude higher than those previously observed. The effects of steric hindrance and electronic configuration on the polymerisation activity of these tantalum and niobium pre‐catalysts were investigated. Spectroscopic studies (1H NMR, 13C NMR and 1H? 1H and 1H? 13C correlations) on the reactions of compounds 4 a / 4 b with either MAO(50) or AlMe3/[CPh3]+[B(C6F5)4]? were consistent with the formation of a diamagnetic cation of the form [L4AlMe2]+ (MAO(50) is the product of the vacuum distillation of commercial MAO at +50 °C and contains only 1 mol % of Al in the form of free AlMe3). In the presence of MAO, this cationic aluminium complex was not capable of initiating the ROMP (ring opening metathesis polymerisation) of norbornene, whereas the 4 a / 4 b systems with MAO(50) were active. A parallel pressure reactor (PPR)‐based homogeneous polymerisation screening by using pre‐catalysts 1 b , 1 c , 2 a , 3 a and 6 a , in combination with MAO, revealed only moderate‐to‐good activities for the homo‐polymerisation of ethylene and the co‐polymerisation of ethylene/1‐hexene. The molecular structures are reported for complexes 1 a – 1 c , 2 b , 5 , 6 a , 6 b, 7 a, 8 a and 8 c .  相似文献   

19.
《化学:亚洲杂志》2017,12(2):239-247
Five bis(quinolylmethyl)‐(1H ‐indolylmethyl)amine (BQIA) compounds, that is, {(quinol‐8‐yl‐CH2)2NCH2(3‐Br‐1H ‐indol‐2‐yl)} ( L1H ) and {[(8‐R3‐quinol‐2‐yl)CH2]2NCH(R2)[3‐R1‐1H ‐indol‐2‐yl]} ( L2–5H ) ( L2H : R1=Br, R2=H, R3=H; L3H : R1=Br, R2=H, R3=i Pr; L4H : R1=H, R2=CH3, R3=i Pr; L5H : R1=H, R2=n Bu, R3=i Pr) were synthesized and used to prepare calcium complexes. The reactions of L1–5H with silylamido calcium precursors (Ca[N(SiMe2R)2]2(THF)2, R=Me or H) at room temperature gave heteroleptic products ( L1, 2 )CaN(SiMe3)2 ( 1 , 2 ), ( L3, 4 )CaN(SiHMe2)2 ( 3 a , 4 a ) and homoleptic complexes ( L3, 5 )2Ca ( D3 , D5 ). NMR and X‐ray analyses proved that these calcium complexes were stabilized through Ca⋅⋅⋅C−Si, Ca⋅⋅⋅H−Si or Ca⋅⋅⋅H−C agostic interactions. Unexpectedly, calcium complexes (( L3–5 )CaN(SiMe3)2) bearing more sterically encumbered ligands of the same type were extremely unstable and underwent C−N bond cleavage processes as a consequence of intramolecular C−H bond activation, leading to the exclusive formation of (E )‐1,2‐bis(8‐isopropylquinol‐2‐yl)ethane.  相似文献   

20.
The synthesis of three 1-(4-trifluoromethylphenyl)-3-methyl-4-R1(C=O)-5-pyrazolone proligands LH (L1H; R1=C6H5: L2H; R1=CH3: L3H; R1=CF3) and their interaction with R3Sn(IV) acceptors (R=Me, Bun, Ph) are reported. When R=Me or Bun, aquo (4-acylpyrazolonate)SnR3(H2O) derivatives are obtained and the anionic donors 4-acylpyrazolonate (L) act in the O–monodentate form. These triorganotin complexes are not stable in chlorohydrocarbon solvents and decompose to R4Sn and bis(4-acyl-5-pyrazolonate)2SnR2. When R=Ph, stable (4-acyl-5-pyrazolonate)SnPh3 derivatives, both in solution and in the solid state, are obtained. The crystal structure of (1-(4-trifluoromethylphenyl)-3-methyl-4-acetylpyrazolon-5-ato)triphenyltin(IV) shows a five-coordinate tin atom in a strongly distorted cis-bipyramidal trigonal environment (axial angle=161.2(2)°) with the acylpyrazolonate donor acting as an asymmetric O2–bidentate species (Sn–O(1)=2.081(6) Å: Sn–O(2)=2.424(5) Å). Electronic effects are responsible for the different behavior shown by these trialkyl and triphenyl derivatives.  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