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 453 毫秒
1.
Reactions of 2,6-dichloro-3-phenyl-, 2,6-dibromo-3-phenyl-, 2,6-dichloro-3-dimethylamino- and 2,6-dibromo-3-dimethylaminopyridine with potassium amide in liquid ammonia were investigated. Whereas 2,6-dichloro-3-phenylpyridine yields 4-amino-2-benzylpyrimidine, from 2,6-dibromo-3-phenylpyridine as a product of a novel ring fission 2-amino-l-cyano-l-phenyl-but-l-en-3-yne was isolated, together with 4-amino-6-bromo-3-phenylpyridine and 2,6-diamino-3-phenylpyridine. It was shown that neither 2-amino-6-bromo-3-phenyl- nor 6-amino-2-bromo-3-phenylpyridine are intermediates in the formation of the 2,6-diamino derivative, as these bromo compounds are transformed in the basic medium into 1,3-dicyano-l-phenylpropene. From both 2,6-di-chloro-3-dimethylamino- and 2,6-dibromo-3-dimethylaminopyridine mixtures are obtained from which only 2-amino-l-cyano-l-dimethylamino-but-l-en-3-yne and 4-amino-6-halogeno-3-dimethylaminopyridine were isolated. Mechanisms for the reactions studied are proposed, i.e. a SN(ANRORC) mechanism for the aminodebromination of 2,6-dibromo-3-phenylpyridine into the corresponding 2,6-diamino compound.  相似文献   

2.
The crucial factor of the reaction of 2,6-di-tert-butylphenol with alkali hydroxides is temperature, depending on which two types of potassium or sodium 2,6-di-tert-butylphenoxides are formed. These types exhibit different catalytic activity in the alkylation of 2,6-di-tert-butylphenol with methyl acrylate. More active forms of 2,6-But 2C6H3OK or 2,6-But 2C6H3ONa are synthesized at temperatures higher than 160 °C and are predominantly the monomers, which dimerize on cooling. The data of 1H NMR, electronic, and IR spectra for the corresponding forms of 2,6-But 2C6H3OK and 2,6-But 2C6H3ONa isolated in the individual state are in agreement with cyclohexadienone structure. In DMSO or DMF, the dimeric forms of 2,6-di-tert-butylphenoxides react with methyl acrylate to form methyl 3-(4-hydroxy-3,5-di-tert-butylphenyl)propionate in 64–92% yield. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2138–2143, December, 2006.  相似文献   

3.
A procedure has been proposed for the synthesis of 2-(cyclopentylsulfanyl)-6-[(1R)-1-(2,6-difluorophenyl) ethyl]-5-methylpyrimidin-4(3H)-one through intermediate (3R)-4,4-dimethyl-2-oxotetrahydrofuran-3-yl (2R)-2-(2,6-difluorophenyl)propanoate which was obtained from prochiral 2-(2,6-difluorophenyl)prop-1-en-1-one generated in situ. The proposed procedure may be regarded as stereoselective route to 6-[(1R)-1-(2,6- difluorophenyl)ethyl]-5-methylpyrimidin-4(3H)-one derivatives.  相似文献   

4.
A new open-chain monoterpene glycoside, anatolioside E ( 1 ), was isolated from the leaves of Viburnum orientale in addition to three known acyclic monoterpene glycosides, betulalbusides A ( 2 ) and B ( 3 ), and 2(E)-2,6-dimethyl-2,7-octadien-1,6-diol-6-O-β-D -glucopyranoside( 4 ). The structure of anatolioside E ( 1 ) was elucidated on the basis of chemical and spectral data as 6-O-[β-D -glucopyransoyl-(1?? → 6?″)-2-(E), 6(R), 2,6-dimethyl-6-hydroxy-2,7-octadienoyl-(1?″ → 2″″)-β-D -glucopyranosyl-(1″″ → 6?)-2-(E), 6(R), 2,6-dimethyl-6-hydroxy-2,7-octadienoyl-(1? → 4″)-α-L -rhamnopyranosyl-(1″″ → 2′)-β-D -glucopyranosyl]linalool.  相似文献   

5.
N-Oxidation of 2-amino-6-chloropurine to the 3-oxide provided a convenient intermediate for the synthesis of 2-amino-6-substituted purine 3-oxides, including the previously unavailable 2,6-diaminopurine 3-oxide. Thiation of the 6-halogen was accompanied by reduction of the N-oxide. The properties of the 1- and 3-oxides of 2,6-diaminopurine are compared.  相似文献   

6.
A new family of t‐butyl substituted chromium(III) chloride complexes ( Cr1 – Cr6 ), bearing 2‐(1‐(2,6‐dibenzhydryl‐4‐t‐butylphenylimino)ethyl)‐6‐(1‐(arylimino)ethyl)pyridine (aryl = 2,6‐Me2C6H3 Cr1 , 2,6‐Et2C6H3 Cr2 , 2,6‐i‐Pr2C6H3 Cr3 , 2,4,6‐Me3C6H2 Cr4 and 2,6‐Et2‐4‐MeC6H2 Cr5 ) or 2,6‐bis(1‐(2,6‐dibenzhydryl‐4‐t‐butylphenylimino)ethyl)pyridine ( Cr6 ), has been synthesized by the reaction of CrCl3·6H2O in good yield with the corresponding ligands ( L1 – L6 ), respectively. The molecular structures of Cr2 and Cr6 were characterized by X‐ray diffraction highlighted a distorted octahedral geometry with the coordinated N,N,N ligand and three bonded chlorides around the metal center. On activation with modified methylaluminoxane or triisobutyl aluminum, most of the chromium precatalysts exhibit good activities toward ethylene polymerization and produce linear polyethylenes with high‐molecular weight. In addition, an in‐depth catalytic evaluation of Cr2 was conducted to investigate how cocatalyst type and amount, reaction temperature, and run time affect the catalytic activities and polymer properties. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019 , 57, 1049–1058  相似文献   

7.
The interaction of 4-oxo-3,4-dihydroquinazolinyl- and benzimidazolylacetonitriles with 2,6-dihalobenzaldehydes leads to 3-(2,6-dihalophenyl)-2-(4-oxo-3,4-dihydro-2-quinazolinyl)acrylonitriles and 2-(1H-benzo[d]imidazol-2-yl)-3-(2,6-dihalophenyl)acrylonitriles respectively. As a result of intramolecular cyclization of these nitriles 4-halo-12-oxo-12H-quino[2,1-b]quinazoline-6-carbonitriles and 4-halobenzo[4,5]imidazo[1,2-a]quinoline-6-carbonitriles respectively are formed.  相似文献   

8.
2,6-Diformylpyridine bis(N4-methylthiosemicarbazone) and bis(N4-dimethylthiosemicarbazone), H22,6Fo4M and H22,6Fo4DM, respectively, and 2,6-diacetylpyridine bis(N4-methylthiosemicarbazone) and bis(N4-dimethylthiosemicarbazone), H22,6Ac4M and H22,6Ac4DM, and their copper(II) and nickel(II) complexes have been synthesized. The 1H-n.m.r. spectra of the free bis(thiosemicarbazones) show that, most often, one of the thiosemicarbazone moieties is hydrogen bonded to the pyridine nitrogen, and in [2H6]-DMSO there is interaction with solvent oxygen. Golden yellow H22,6Ac4DM has a bifurcated hydrogen bonding interaction by one of the thiosemicarbazone moieties resulting in conjugation. Coordination to copper(II) and nickel(II) centers is via the pyridine nitrogen, amine nitrogen and thiolato sulfur and most of the complexes formed are polynuclear with thiosemicarbazone moieties from the same ligand coordinating to different metal centers.  相似文献   

9.
Fluoride abstraction from bis-m-terphenylelement fluorides (2,6-Mes2C6H3)2EF (E=P, As) generated the highly reactive phosphenium ion [(2,6-Mes2C6H3)2P]+ and the arsenium ion [(2,6-Mes2C6H3)2As]+, which immediately underwent intramolecular electrophilic substitution and formation of an 1,2,4-trimethyl-6-mesityl-5-m-terphenyl-benzo[b]phospholium ion and an 1,2,4-trimethyl-6-mesityl-5-m-terphenyl-benzo[b]arsolium ion, respectively. The formation of the latter involved a methyl group migration from the ortho-position of a flanking mesityl group to the meta-position. This reactivity of [(2,6-Mes2C6H3)2E]+ (E=P, As) is in sharp contrast to the related stibenium ion [(2,6-Mes2C6H3)2Sb]+ and bismuthenium ion [(2,6-Mes2C6H3)2Bi]+, which have been recently isolated and fully characterized (Angew. Chem. Int. Ed. 2018, 57 , 10080–10084). On the basis of DFT calculations, a mechanism for the rearrangement of the phosphenium and arsenium ions into the phospholium and arsolium ions is proposed, which is not feasible for the stibenium and bismuthenium ions.  相似文献   

10.
Three new acacic acid derivatives, named coriariosides C, D, and E ( 1–3 ) were isolated from the roots of Albizia coriaria. Their structures were elucidated on the basis of extensive 1D‐ and 2D‐NMR studies and mass spectrometry as 3‐O‐[β‐D ‐xylopyranosyl‐(1 → 2)‐β‐D ‐fucopyranosyl‐(1 → 6)‐2‐(acetamido)‐2‐deoxy‐β‐D ‐glucopyranosyl]‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐ 6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐β‐D ‐quinovopyranosyl}‐2,6‐dimethylocta‐2,7‐dienoyl}acacic acid 28‐O‐β‐D ‐xylopyranosyl‐(1 → 4)‐α‐L ‐rhamnopyranosyl‐(1 → 2)‐β‐D ‐glucopyranosyl ester ( 1 ), 3‐O‐{β‐D ‐fucopyranosyl‐(1 → 6)‐[β‐D ‐glucopyranosyl‐(1 → 2)]‐β‐D ‐glucopyranosyl}‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐β‐D ‐quinovopyranosyl}‐2,6‐dimethylocta‐2,7‐dienoyl}acacic acid 28‐O‐α‐L ‐rhamno pyranosyl‐(1 → 2)‐β‐D ‐glucopyranosyl ester ( 2 ), and 3‐O‐[β‐D ‐fucopyranosyl‐(1 → 6)‐β‐D ‐glucopyranosyl]‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl)‐β‐D ‐quinovopyranosyl]octa‐2,7‐dienoyl}acacic acid 28‐O‐β‐D ‐glucopyranosyl ester ( 3 ). Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

11.
Reaction of phosphorus oxychloride with 2,6-dimethylpyridine N-oxide hydrochloride ( 1 ) gave a mixture of 2-(chloromethyl)-6-methylpyridine ( 2 ) and 4-chloro-2,6-dimethylpyridine ( 3 ). Treatment of this mixture with triethylamine converted 2 to the quaternary salt 4 which was separated by water extraction leaving 3 which was subsequently reacted with trimethylstannyl sodium to yield 2,6-dimethyl-4-(trimethylstannyl)pyridine ( 6 ).  相似文献   

12.
Resolution and Determination of the Absolute Configuration of 2,6-Disubstituted Bicyclo[3.3.1]nonanes (±)-endo, endo-Bicyclo [3.3.1]nonane-2,6-diol was resolved via diastereomeric camphanic acid esters. Conversion of the (+)-enantiomer 2 via (+)- 5 and (+)- 6 as key intermediates gave (+) methyl 3-(3-oxocyclohexyl)-propionate ( 7 ) which independently could be prepared also from the known (+)-(R)-3-oxo-cyclohexane-carboxylic acid ( 8 ). These chemical correlations establish the absolute configuration of (+) -2 , (+) -5 and (+) -6 as well as that of (+)-bicyclo [3.3.1]nonane-2,6-dione ( 1 ) obtained by oxidation of (+) -2 . The chiroptical properties of 1 and 6 are discussed.  相似文献   

13.
Treatment of the chlorides (L2,6‐iPr2Ph)2LnCl (L2,6‐iPr2Ph = [(2,6‐iPr2C6H3)NC(Me)CHC(Me)N(C6H5)]?) with 1 equiv. of NaNH(2,6‐iPr2C6H3) afforded the monoamides (L2,6‐iPr2Ph)2LnNH(2,6‐iPr2C6H3) (Ln = Y ( 1 ), Yb ( 2 )) in good yields. Anhydrous LnCl3 reacted with 2 equiv. of NaL2,6‐iPr2Ph in THF, followed by treatment with 1 equiv. of NaNH(2,6‐iPr2C6H3), giving the analogues (L2,6‐iPr2Ph)2LnNH(2,6‐iPr2C6H3) (Ln = Sm ( 3 ), Nd ( 4 )). Two monoamido complexes stabilized by two L2‐Me ligands, (L2‐Me)2LnNH(2,6‐iPr2C6H3) (L2‐Me = [N(2‐MeC6H4)C(Me)]2CH)?; Ln = Y ( 5 ), Yb ( 6 )), were also synthesized by the latter route. Complexes 1 , 2 , 3 , 4 , 5 , 6 were fully characterized, including X‐ray crystal structure analyses. Complexes 1 , 2 , 3 , 4 , 5 , 6 are isostructural. The central metal in each complex is ligated by two β‐diketiminato ligands and one amido group in a distorted trigonal bipyramid. All the complexes were found to be highly active in the ring‐opening polymerization of L‐lactide (L‐LA) and ε‐caprolactone (ε‐CL) to give polymers with relatively narrow molar mass distributions. The activity depends on both the central metal and the ligand (Yb < Y < Sm ≈ Nd and L2‐Me < L2,6‐iPr2Ph). Remarkably, the binary 3/benzyl alcohol (BnOH) system exhibited a striking ‘immortal’ nature and proved able to quantitatively convert 5000 equiv. of L‐LA with up to 100 equiv. of BnOH per metal initiator. All the resulting PLAs showed monomodal, narrow distributions (Mw/Mn = 1.06 ? 1.08), with molar mass (Mn) decreasing proportionally with an increasing amount of BnOH. The binary 4/BnOH system also exhibited an ‘immortal’ nature in the polymerization of ε‐CL in toluene. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

14.
Five new acyclic monoterpene glycosides 1 – 5 were isolated from the leaves of Viburnum orientale (Caprifoliaceae). Anatolioside ( 1 ) is a monoterpene diglycoside and its structure was elucidated as linalo-6-yl 2′-O-(α-L -rhamnopyranosyl)β-D -glucopyranoside (arbitrary numbering of linalool moiety). Compounds 2 – 5 are all derivatives of 1 , containing additional monoterpene and sugar units, connected by ester and glycoside bonds. Their structures were established as linalo-6-yl O-[(2E,6R)-6-hydroxy-2, 6-dimethylocta-2,7-dienoyl]-(1? → 4″)-O-α-L -rhamnopyranosyl-(1″? → 2″″)-β-D -glucopyranoside ( = anatolioside A; 2 ), linalo-6-yl O-β-D -glucopyranosyl-(1? → 6?)-O-[(2E,6R)-6-hydroxy-2,6-dimethylocta-2,7-dienoyl]-(1? → 4″)-O-α-L -rhamnopyranosyl-(1″ → 2′)–β-D -glucopyranoside ( = anatolioside B; 3 ), linalo-6-yl O-β-D ribo-hexopyranos-3-ulosyl-(1′? → 6?)-O-[(2E,6R)-6-hydroxy-2,6-dimethylocta-2,7-dienoyl]-(1? → 4″)-O-α-L -rhamnopyranosyl-(1″ → 2′)-β-D -glucopyranoside ( = anatolioside C; 4 ) and linalo-6-yl O-[(2E, 6R)-6-hydroxy-2,6-dimethylocta-2,7-dienoyl]-(1″? → 2″″)-O-β-D -glucopyranosly-(1″″ → 6?)-O-[(2E,6R)-6-hydroxy-2,6-dimethylocta-2,7-dienoyl]-(1? → 4″)-O-α-L -rhamnopyranosyl(1″ → 2′)-β-D -glucopyranoside ( = anatolioside D ; 5 ). The structure determinations were based on spectroscopic and chemical methods (acid and alkaline hydrolysis, acetylation and methylation).  相似文献   

15.
The structures of cocrystals of 2,6‐dichlorophenol with 2,4‐diamino‐6‐methyl‐1,3,5‐triazine, C6H4Cl2O·C4H7N5, (III), and 2,6‐dichloroaniline with 2,6‐diaminopyrimidin‐4(3H)‐one and N,N‐dimethylacetamide, C6H5Cl2N·C4H6N4O·C4H9NO, (V), plus three new pseudopolymorphs of their coformers, namely 2,4‐diamino‐6‐methyl‐1,3,5‐triazine–N,N‐dimethylacetamide (1/1), C4H7N5·C4H9NO, (I), 2,4‐diamino‐6‐methyl‐1,3,5‐triazine–N‐methylpyrrolidin‐2‐one (1/1), C4H7N5·C5H9NO, (II), and 6‐aminoisocytosine–N‐methylpyrrolidin‐2‐one (1/1), C4H6N4O·C5H9NO, (IV), are reported. Both 2,6‐dichlorophenol and 2,6‐dichloroaniline are capable of forming definite synthon motifs, which usually lead to either two‐ or three‐dimensional crystal‐packing arrangements. Thus, the two isomorphous pseudopolymorphs of 2,4‐diamino‐6‐methyl‐1,3,5‐triazine, i.e. (I) and (II), form a three‐dimensional network, while the N‐methylpyrrolidin‐2‐one solvate of 6‐aminoisocytosine, i.e. (IV), displays two‐dimensional layers. On the basis of these results, attempts to cocrystallize 2,6‐dichlorophenol with 2,4‐diamino‐6‐methyl‐1,3,5‐triazine, (III), and 2,6‐dichloroaniline with 6‐aminoisocytosine, (V), yielded two‐dimensional networks, whereby in cocrystal (III) the overall structure is a consequence of the interaction between the two compounds. By comparison, cocrystal–solvate (V) is mainly built by 6‐aminoisocytosine forming layers, with 2,6‐dichloroaniline and the solvent molecules arranged between the layers.  相似文献   

16.
Novel Silanes with Sterically Demanding Aryl Substituents A series of novel m‐terphenylsilanes was synthesized and fully characterized. Halogenation of Si(C6H3‐2,6‐Trip2)H3 ( 1 , Trip = C6H2‐2,4,6‐iPr3) with ICl and BI3 afforded the m‐terphenylmonochlorosilane Si(C6H3‐2,6‐Trip2)H2Cl ( 2 ) and the m‐terphenyldiiodosilane Si(C6H3‐2,6‐Trip2)HI2 ( 3 ), respectively. The chiral phosphanylmethylsilane Si(C6H3‐2,6‐Trip2)HCl(CH2PMe2) ( 6 ) was obtained by metathetical exchange of Si(C6H3‐2,6‐Trip2)HCl2 ( 4 ) with the Grignard reagent Me2PCH2MgCl ( 5 ). Similarly, metathetical exchange of SiBr4 with 0.5 equiv. of {Li(C6H3‐2,6‐Mes)}2 ( 7 ) yielded the m‐terphenyltribromsilane Si(C6H3‐2,6‐Mes2)Br3 ( 8 ). A Pd(II) catalyzed reaction of MesSiH3 ( 10 ) with 2‐iodopropane afforded the triiodosilane MesSiI3 ( 11 ) bearing the sterically less demanding mesityl substituent. The silanes were fully characterized and the molecular structures of compounds 2 and 11 were determined by single‐crystal X‐ray diffraction.  相似文献   

17.
Zusammenfassung Die Konfiguration der 10 stereoisomeren 2,6-Dimethyl-, 2,6-Diäthyl- und 2-Methyl-6-äthylcyclohexylamine wird bestimmt. Die Zuordnung erfolgt durch Vergleich der experimentell erhaltenen NMR-Spektren mit den auf Grund energetischer Konformationsbetrachtungen erwarteten-Signalen. Die Messungen werden an den konfigurationsgleichen Nitroverbindungen vorgenommen.
Summary The configurations of the 10 stereoisomeric 2,6-dimethyl-, 2,6-diethyl-, and 2-methyl-6-ethylcyclohexylamines are determined. The determination is carried out by comparing the measured NMR-spectra of the nitro compounds of identical configuration with the expected-signals deduced by conformational considerations.
  相似文献   

18.
A series of new titanium(IV) complexes with o‐metalated arylimine and/or cis‐9,10‐dihydrophenanthrenediamide ligands, [o‐C6H4(CH?NR)TiCl3] (R=2,6‐iPr2C6H3 ( 3 a ), 2,6‐Me2C6H3 ( 3 b ), tBu ( 3 c )), [cis‐9,10‐PhenH2(NR)2TiCl2] (PhenH2=9,10‐dihydrophenanthrene; R=2,6‐iPr2C6H3 ( 4 a ), 2,6‐Me2C6H3 ( 4 b ), tBu ( 4 c )), [{cis‐9,10‐PhenH2(NR)2}{o‐C6H4(HC?NR)}TiCl] (R=2,6‐iPr2C6H3 ( 5 a ), 2,6‐Me2C6H3 ( 5 b ), tBu ( 5 c )), have been synthesised from the reactions of TiCl4 with o‐C6H4(CH?NR)Li (R=2,6‐iPr2C6H3, 2,6‐Me2C6H3, tBu). Complexes 4 and 5 were formed unexpectedly from the reactions of TiCl4 with two or three equivalents of the corresponding o‐C6H4(CH?NR)Li followed by sequential intramolecular C? C bond‐forming reductive elimination and oxidative coupling reactions. Attempts to isolate the intermediates, [{o‐C6H4(CH?NR)}2TiCl2] ( 2 ), were unsuccessful. All complexes were characterised by 1H and 13C NMR spectroscopy, and the molecular structures of 3 a , 4 a – c , 5 a , and 5 c were determined by X‐ray crystallography.  相似文献   

19.
[(1-Amino-6-hydroxy-2(1H)-pyrimidinylidene)hydrazone]butanedioic acid dimethyl esters 3 , formed from 3-amino-2-hydrazino-4(3H)-pyrimidinones and dimethyl acetylenedicarboxylate in acetic acid at room temperature, underwent a facile, thermal rearrangement to 1-amino-2,6-dihydro-2,6-dioxo-1H-pyrimido-[1,2-b]-[1,2,4]triazine-3-acetic acid methyl esters 6 in hot acetic acid.  相似文献   

20.
Diastereoisomeric couples of (1R,5R,6RS)-2,6-dimethyl-7-oxabicyclo[3.2.1]oct-2-en-6-ylmethanols and their epoxide precursors, (5R,2′RS)-2-methyl-5-(2-methyloxiran-2-yl)cyclohex-2-en-1-ols, in the presence of triethylsilyl trifluoromethanesulfonate underwent [3.2.1]→[3.3.1] skeletal rearrangement with formation of scalemic mixtures of (1R,5R,6R)- and (1R,5R,6S)-2,6-dimethyl-6-triethylsiloxy-8-oxabicyclo[3.3.1]non-2-enes; the (6R)-stereoisomer was isolated as individual substance.  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