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 reductive amination of (?)‐2‐carene‐3‐aldehyde, prepared in two steps from (?)‐perillaldehyde, furnished 2‐carene‐based allylamines. tert‐Butyloxycarbonyl (Boc) or carbobenzyloxy (Cbz) protection of the resulting amines, followed by stereoselective dihydroxylation in highly stereospecific reactions with OsO4 and subsequent deprotection, resulted in N‐benzylaminodiols, which were transformed to primary and tertiary aminodiols. The reactions of the N‐benzyl‐ and N‐(1‐phenylethyl)‐substituted derivatives with formaldehyde led to highly regioselective ring closure, resulting in carane‐fused 1,3‐oxazines. The aminodiols and their 1,3‐oxazine derivatives were applied as chiral catalysts in the enantioselective addition of diethylzinc to aldehydes. The best (R) enantioselectivity was observed in the case of the N‐((R)‐1‐phenylethyl)‐substituted aminodiol, whereas the opposite chiral direction was preferred when the 1,3‐oxazines were applied. Through the use of molecular modelling at an ab initio level, this phenomenon was interpreted in terms of competing reaction pathways. Molecular modelling at the RHF/LANL2DZ level of theory was successfully applied for a mechanism‐based interpretation of the stereochemical outcome of the reactions leading to the development of further 1,3‐oxazine‐based ligands, which display excellent (S) enantioselectivity (95 and 98 % ee) in the examined transformation.  相似文献   

2.
Chlorination of 2-thiophenesulfonamide gave unstable N,N-dichloro-2-thiophenesulfonamide which was brought into reactions with 1,2-polyhaloethenes. The condensation of 2-thiophenesulfonamide with trichloroacetaldehyde afforded N-(2,2,2-trichloro-1-hydroxyethyl)-2-thiophenesulfonamide which reacted with benzene, toluene, 2-chlorothiophene, and phenol to form the corresponding N-(1-aryl-2,2,2-trichloroethyl)-2-thiophenesulfonamides. Under more severe conditions, the latter were converted into 1,1-diaryl-2,2,2-trichloroethanes. The reaction of N-(2,2,2-trichloro-1-hydroxyethyl)-2-thiophenesulfonamide with substituted arenes, including phenol, was regioselective: only the corresponding para-substituted products were obtained. Hydrolysis of N-[2,2,2-trichloro-1-(4-tolyl)ethyl]-2-thiophenesulfonamide yielded N-(2-thienylsulfonyl)-2-(4-tolyl)glycine.  相似文献   

3.
Many observations prove that a number of silylation reactions of a trialkylsilyl halide-uncharged base system occur with the transient formation of a 1:1 tetrahedral silicon ionic complex of the silyl halide with the base. Some catalytic processes of phosphorylation of protonic substrates with tricoordinate phosphorus halides in mixture with an uncharged base show similar features to these silylation reactions, implying that a similar mechanism may operate. It was demonstrated that Ph2PCl phosphorylates t-BuOH faster under catalysis with 4-N,N-dimethylamino pyridine or N-methylimidazole than in the presence of Et3N by a factor of 400 and 33, respectively. The catalytic phosphorylation process exhibits a very low activation energy and a high negative value of entropy of activation. The interaction of the uncharged bases with model tricoordinate phosphorus halides was demonstrated to lead to the formation of ionic 1:1 complexes without changing the coordination number of phosphorus, in full analogy to the silyl halide complex formation. Finally, the interaction of phosphorous tris(dimethylamide) with a silyl iodide and a phosphorous iodide results in both cases in the formation of the ionic 1:1 complex, which also leads to analogous reactions of exchange of the amide group with iodide. These close similarities imply that some phosphorylation reactions with tricoordinate phosphorus halides catalyzed with uncharged bases occur via a tricoordinate phosphorus cation intermediate.  相似文献   

4.
Bridged N,N′-substituted bisbenzimidazolium bromide salts (L1, L2, and L3) were synthesized and fully characterized. Reactions of palladium acetate with L1, L2, and L3 afforded corresponding new bridged bis(N-heterocyclic carbene)palladium(II) complexes (C1, C2, and C3) in high yields. The X-ray structure of complex C1 showed that the Pd(II) ion is bonded to the two carbon atoms of the bis(N-heterocyclic carbene) and two bromido ligands are in the cis position, resulting in a distorted square planar geometry. The three Pd(NHC)2Br2 complexes C1, C2, and C3 were evaluated in carbonylative Suzuki–Miyaura coupling reactions of aryl boronic acids with aryl halides and displayed high catalytic activity with low catalyst loading. The coupling reactions of aryl bromides were selective towards the carbonylation product at higher carbon monoxide pressure.  相似文献   

5.
The reactions of oxybis(diacetoxyborane) with the aldimines, N-(2-hydroxyethyl) salicylaldimine N-(2-hydroxy-1-propyl) salicylaldimine, N-(3-hydroxy-1-propyl) salicylaldimine, N-(o-hydroxyphenyl) salicylaldimine. N-(m-hydroxyphenyl) salicylaldimine, N(2-hydroxyethyl)-2-hydroxy-1-naphthaldimine and N(2-hydroxy-1-propyl) 2-hydroxy-1-naphthaldimine have been carried out in 1: 1 and 1: 2 molar ratios. All the compounds except those derived from N-(3-hydroxy-1-propyl) salicylaldimine have been found to be sparingly soluble in benzene and nonelectrolytes in anhydrous DMF. The newly synthesized derivatives have been characterized by elemental analysis, molecular weight determinations and infrared, proton magnetic resonance, ultraviolet, visible and 11B nuclear magnetic resonance spectral studies.  相似文献   

6.
Aroylation of 5,6,8,8a,13,14,16,16a-octahydro[1,2,4,5]tetrazino[6,1-a:3,4-a′]diisoquinoline or 1,3,4,8b-tetrahydro[1,2]diazireno[3,1-a]isoquinoline, as well as reactions of 2-(2-bromoethyl)benzaldehyde with aroylhydrazines followed by treatment with triethylamine, led to the formation of stable azomethine imines, aroyl(3,4-dihydroisoquinolinium-2-yl)azanides. 1,3-Dipolar cycloaddition of the latter to N-mesitylmaleimide was stereoselective: the ratio of the trans- and cis-adducts was ∼(3–7): 1, the former prevailing. The reactions with N-arylmaleimides having no ortho-substituents in the aryl group gave the corresponding cis-adducts as the major products [trans/cis ratio ∼1: (2.5–10)].  相似文献   

7.
Bis-aminimide compounds [bis-N, N,-dimethyl-N,-(2-hydroxypropyl)-amine-N,′-adipimide ( 1 ) and bis-trimethylamine adipimide ( 2 )] were found to exhibit different thermal decomposition behavior and polymerization efficiency for an epoxide (phenyl glycidyl ether, PGE). The thermal decomposition rate of 1 was much higher than that of 2. It seemed that hydrogen bonding enhanced the decomposition rate. Compound 1 was thermolyzed to give a diisocyanate and a tertiary aminoalcohol, which subsequently reacted with each other to give a urethane. When 2 was heated, the isocyanate generated from 2 remained unreacted. PGE reacted with thoseaminimides to give different products, depending on their thermolyzed products. Mixtures of diisocyanate, tertiary amine, and PGE were used in the model reactions, and the thermal reaction between the expected decomposition products of aminimides was investigated in the presence and absence of PGE. The amount of high molecular weight fraction in PGE + 2 is greater than that of PGE + 1 . In the former, the free isocyanate groups may act as a chain extender to give higher molecular weight fractions.  相似文献   

8.
Triethylgermylation of sulfacetamide occurs on the sulfonamido nitrogen in competition with the 1,2 addition of the starting triethylgermyl dimethylamine on the carbonyl group. Thermal decomposition in the presence of dimethylamine yields N‐triethylgermylsulfanilamide. Stable 1:1 sulfacetamide–DBU and 1:1 sulfacetamide–Et3N complexes were isolated and fully characterized in the course of dehydrochlorination reactions. o‐Sulfonamidophenylamine yields N,N′‐bis‐triethylgermylated derivatives, whereas o‐(N,N‐dimethylsulfonamido)phenylamine leads to monogermylated compounds. The N‐dimethylaminodimesitylgermyl derivative is thermally stable. Dehydrohalogenation of the N‐dimesitylfluorogermyl compound leads to the thermally stable but water sensitive N‐[o‐(N′,N′‐dimethylsulfonamido)phenyl]‐N‐dimesitylgermaimine. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

9.
1‐(2‐N,N‐Dimethylaminoethyl)‐2,3,4,5‐tetramethylcyclopentadienyl‐chromium dichloride ( 1 ), (2‐N,N‐dimethylaminoethyl)cyclopentadienylchromium dichloride ( 6 ), and (2‐N,N‐dimethylaminoethyl)indenylchromium dichloride ( 7 ) in the presence of modified methylaluminoxane exhibit high catalytic activities for the polymerization of ethylene with random copolymerizations of ethylene with propylene, ethylene with 1‐hexene, and propylene with 1‐hexene. These initiators conduct polymerizations to give high molecular weight polymers with low polydispersities. However, the stereoregularities are very poor in these reactions. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 2759–2771, 2002  相似文献   

10.
Addition reactions of thioamide dianions that were derived from N‐arylmethyl thioamides to imines and aziridines were carried out. The reactions of imines gave the addition products of N‐thioacyl‐1,2‐diamines in a highly diastereoselective manner in good‐to‐excellent yields. The diastereomeric purity of these N‐thioacyl‐1,2‐diamines could be enriched by simple recrystallization. The reduction of N‐thioacyl‐1,2‐diamines with LiAlH4 gave their corresponding 1,2‐diamines in moderate‐to‐good yields with retention of their stereochemistry. The oxidative‐desulfurization/cyclization of an N‐thioacyl‐1,2‐diamine in CuCl2/O2 and I2/pyridine systems gave the cyclized product in moderate yield and the trans isomer was obtained as the sole product. On the other hand, a similar cyclization reaction with antiformin (aq. NaClO) as an oxidant gave the cis isomer as the major product. The reactions of N‐tosylaziridines gave the addition products of N‐thioacyl‐1,3‐diamines with low diastereoselectivity but high regioselectivity and in good‐to‐excellent yields. The use of AlMe3 as an additive improved the efficiency and regioselectivity of the reaction. The stereochemistry of the obtained products was determined by X‐ray diffraction.  相似文献   

11.
A number of new and interesting 2-amino-4-(N-substituted)imino-4,5-dihydrothiazoles were synthesized by reacting thiourea (or thiourea hydrochloride) with N-alkyl- or N,N-dialkyl-N′-p-toluenesulfonyl-α-chloroacetamidines, where the N,N-alkyl groups were ethyl, cyclohexyl, benzyl, β-phenethyl, (3,5-dimethyl-1-adamantyl)-methyl, as well as N,N-dimethyl- and N,N-pentamethylene. Reactions of N-alkyl-N-p-toluenesulfonyl-2-chloroacetamidines (substituents being N-ethyl, N-benzyl and N,N-dimethyl) with thiourea hydrochloride in hot 2-propanol furnished 2-amino-4-(p-toluenesulfonyl)imino-4,5-dihydrothiazole (in 51, 60 and 65% yields, respectively) and the corresponding amine hydrochloride. In hot acetone or butanone, the reactions of these N-sulfonyl-2-chloroacetamidines with excess thiourea provided 2-amino-4-N-(alkyl or N,N-dialkyl)imminium-4,5-dihydrothiazole chlorides in 25–80% yield. The by-product from these reactions was p-toluenesulfonamide. The structures of the products were established by chemical transformations and spectral methods (nmr and mass spectra).  相似文献   

12.
The title compounds were prepared by aldol reaction of anisaldehyde and the respective N,N‐dibenzyl glycinates. Deprotection of the nitrogen atom with Pearlman’s catalyst delivered the unprotected β‐hydroxytyrosine esters, which were further N‐protected as N,N‐phthaloyl (Phth) and N‐fluorenylmethylcarbonyloxy (Fmoc) derivatives. The Friedel–Crafts reaction with various arenes was studied employing these alcohols as electrophiles. It turned out that the facial diastereoselectivitiy depends on the nitrogen protecting group and on the ester group. The unprotected substrates (NH2) gave preferentially syn‐products but the anti‐selectivity increased when going from NHFmoc over NPhth to NBn2. If the ester substituent was varied the syn‐preference increased in the order Me <Et <iPr. The reactions were shown to be fully stereoconvergent and proceeded under kinetic product control. A model is suggested to explain the facial diastereoselectivity based on a conformationally locked benzylic cation intermediate. The reactions are preparatively useful for the N‐unprotected isopropyl ester, which gave Friedel–Crafts alkylation products with good syn‐selectivity (anti/syn=21:79 to 7:93), and for the N,N‐dibenzyl‐protected methyl ester, which led preferentially to anti‐products (anti/syn=80:20 to >95:5). Upon acetylation of the latter compound to the respective acetate, Bi(OTf)3‐catalyzed alkylation reactions became possible, in which silyl enol ethers served as nucleophiles. The respective alkylation products were obtained in high yield and with excellent anti‐selectivitiy (anti/syn≥95:5).  相似文献   

13.
The reactions of two diaminotriazine ligands 2,4‐diamino‐6‐(2‐pyridyl)‐1,3,5‐triazine (2‐pydaT) and 6‐phenyl‐2,4‐diamino‐1,3,5‐triazine (PhdaT) with ruthenium–arene precursors led to a new family of ruthenium(II) compounds that were spectroscopically characterized. Four of the complexes were cationic, with the general formula [(η6‐arene)Ru(κ2N,N‐2‐pydaT)Cl]X (X=BF4, TsO; arene=p‐cymene: 1.BF4 , 1.TsO arene=benzene: 2.BF4 , 2.TsO ). The neutral cyclometalated complex [(η6p‐cymene)Ru(κ2C,N‐PhdaT*)Cl] ( 3 ) was also isolated. The structures of complexes 2.BF4 and 3.H2O were determined by X‐ray diffraction. Complex 1.BF4 underwent a partial reversible‐aquation process in water. UV/Vis and NMR spectroscopic measurements showed that the reaction was hindered by the addition of NaCl and was pH‐controlled in acidic solution. At pH 7.0 (sodium cacodylate) Ru–Cl complex 1.BF4 was the only species present in solution, even at low ionic strength. However, in alkaline medium (KOH), complex 1.BF4 underwent basic hydrolysis to afford a Ru–OH complex ( 5 ). Fluorimetric studies revealed that the interaction of complex 1.BF4 with DNA was not straightforward; instead, its main features were closely linked to ionic strength and to the [DNA]/complex ratio. The bifunctional complex 1.BF4 was capable of interacting concurrently through both its p‐cymene and 2‐pydaT groups. Cytotoxicity and genotoxicity studies showed that, contrary to the expected behavior, the complex species was biologically inactive; the formation of a Ru–OH complex could be responsible for such behavior.  相似文献   

14.
N-Pyridylmethyl-substituted Ephedra derivatives were synthesized by either direct alkylation or reductive alkylation of (1R,2S)-norephedrine, (1S,2S)-pseudo norephedrine, and (1R,2S)-ephedrine. These derivatives were then employed in asymmetric addition reactions with diethylzinc and aldehydes and diphenylphosphinoylimines. The use of the diastereomers from the Ephedra family allowed for a systematic evaluation of the contribution of the N-pyridylmethyl.  相似文献   

15.
The syntheses of several aminoalkylferrocenes by the solid-state reactions of trimethylammoniomethylferrocene iodides and (S)-(–)-(1-trimethylammonio)ethylferrocene with primary amines are described. New ferrocenylalkylamines were synthesized: 1-(4-toluidino)ethylferrocene, 1-(2-pyridylamino)ethylferrocene, 2-pyridylaminomethylferrocene, and N,N-bis(ferrocenylmethyl)-2-pyridylamine.  相似文献   

16.
Kinetic results are reported for the reactions of 4‐nitrophenyl‐2,4,6‐trinitrophenyl ether 3 with aniline and N‐methylaniline in dimethyl sulphoxide, acetonitrile, methanol, and benzene. The reactions gave the expected 2,4,6‐trinitrodiphenylamine and were base catalyzed in all the solvents. Both nucleophiles showed the same kinetic pattern under the same reaction conditions but aniline was found to be considerably more reactive than N‐methylaniline. The greater catalytic efficiency of aniline over N‐methylaniline is consistent with the proton transfer mechanism of the base‐catalyzed step. Dichotomy of amine effects in aromatic substitution (SNAr) reactions is discussed. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 188–196, 2004  相似文献   

17.
The reactions of phenanthrene 9,10-imine ( 1 ) with aromatic aldehydes, benzoic acids and acetylenedi-carboxylic esters were investigated. The aldehydes were shown to give 1-[N-(arylmethylidene)-9-phenanthreneamine-10-yl]-1a,9b-dihydrophenanthro[9,10-b]azirine 2. The ‘dimeric’ structure of these products was established by X-ray diffraction analysis. The carboxylic acids proved to form in the presence of dicyclohexylcarbodiimide, N-aroylphenanthrene 9,10-imines 7 , that readily undergo rearrangement to N-aroyl-9-phenanthrenamines 8. Esters of acetylenedicarboxylic acid gave the corresponding esters of (Z)-2-(1a,9b-dihydrophenanthro[9,10-b]azirine-1-yl)-2-butendioic acid 10 .  相似文献   

18.
The reactions of a series of anilines with phenyl 2,4‐dinitrophenyl ( 1 ), 4‐nitrophenyl 2,4‐dinitrophenyl ( 2 ), and bis(2,4‐dinitrophenyl) ( 3 ) carbonates are subjected to a kinetic investigation in 44 wt% ethanol–water, at 25.0 ± 0.1°C and an ionic strength of 0.2 M. Under amine excess pseudo‐first‐order rate coefficients (kobs) are obtained. Plots of kobs against free amine concentration at constant pH are linear, with slopes kN. The Brønsted plots (log kN vs. anilinium pKa) for the anilinolysis of 1 – 3 are linear, with slope (β) values of 0.52, 0.61, and 0.63, respectively. The values of these slopes and other considerations suggest that these reactions are ruled by a concerted mechanism. For these reactions, the kN values follow the reactivity sequence: 3 > 2 > 1 . Namely, the reactivity increases as the number of nitro groups attached to the nonleaving group increases. Comparison of the reactions of this work with the stepwise pyridinolysis of carbonates 1 – 3 indicates that the zwitterionic tetrahedral intermediate (T±) formed in the pyridinolysis reactions is destabilized by the change of its pyridino moiety by an isobasic anilino group. This is attributed to the superior leaving ability from the T± intermediate of anilines, relative to isobasic pyridines, which destabilize kinetically this intermediate. The kN values for the anilinolysis of carbonates 1 – 3 are similar to those found in the reactions of these carbonates with secondary alicyclic amines. With the kinetic data for the anilinolysis of the title substrates and 4‐methylphenyl and 4‐chlorophenyl 2,4‐dinitrophenyl carbonates, a multiparametric equation is derived for log kN as a function of the pKa of the conjugate acids of anilines and nonleaving groups. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 191–197, 2011  相似文献   

19.
The reactions of enaminones with dimethyl diazomalonate were investigated in the presence of copper(II) acetylacetonate. From the reaction of (E)‐3‐[methyl(phenyl)amino]‐1‐phenylprop‐2‐en‐1‐one ( 6c ), dimethyl 2‐[methyl(phenyl)amino]‐4‐oxonaphthalene‐1,1‐(4H)‐dicarboxylate, was unexpectedly obtained as the major product. Quinoline derivatives were formed as the major products in the case of N‐methyl‐p‐anisidino and N‐methyl‐p‐toluidino enaminones. The reactions of acetyl enaminones were also realized, and quinoline derivatives were isolated as the major products. 3H‐ and 5H‐dihydrofurans were also formed as side products in these reactions. These results differ from those reported earlier on the reactions of tertiary enaminones with carbenes/metal carbenes.  相似文献   

20.
The elimination kinetics of ethyl N‐benzyl‐N‐cyclopropylcarbamate and ethyl diphenylcarbamate were investigated over the temperature range of 349.9–440.0°C and the pressure range of 31–106 Torr. These reactions have been found to be homogeneous, unimolecular, and obey a first‐order rate law. The products are ethylene, carbon monoxide, and the corresponding secondary amine. The rate coefficient is expressed by the following Arrhenius equations: For ethyl N‐benzyl‐N‐cyclopropylcarbamate log k1 (s?1) = (12.94 ± 0.09) ? (198.5 ± 0.9) kJ mol?1 (2.303RT)?1 For ethyl diphenylcarbamate log k1 (s?1) = (12.91 ± 0.18) ? (208.2 ± 2.4) kJ mol?1 (2.303RT)?1 The presence of phenyl and bulky groups at the nitrogen atom of the ethylcarbamate showed a decrease in the rate of elimination. Steric factor may be operating during the process of decomposition of these substrates. These reactions appear to undergo a semipolar six‐membered cyclic transition type of mechanism.© 2001 John Wiley & Sons, Inc. Int J Chem Kinet 34: 67–71, 2002  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