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.
Summary.  Reactions of cyanomethanesulfonamides with aromatic aldehydes in the presence of AcOH and piperidine produced the addition products, the 1-cyano-2-arylethenesulfonamides, whereas reactions with benzonitrile yielded the 2-amino-1-cyano-2-phenylethenesulfonamides only when done in THF with BuLi. No addition products were isolated from the analogue reactions with 2-hydroxybenzaldehyde (salicylaldehyde). Instead, we obtained 2-imino-2H-chromene-3-sulfonamides with good to excellent yields. These 2H-chromene derivatives allowed a number of transformations, from which the reactions with orthoformates opened an approach to the hitherto unknown benzopyrano[3,2-e] [1,2,4]thiadiazine ring system.  相似文献   

2.
Abstract

DAST-assisted rearrangement of 3-O-allyl-4-O-benzyl-α-l-rhamnopyranosyl azide followed by treatment of the generated fluorides with ethanethiol and BF3·OEt2 gave glycosyl donor ethyl 3-O-allyl-2-azido-4-O-benzyl-2,6-dideoxy-1-thio-α/β-l-glucopyranoside. Stereoselective glycosylation of methyl 4,6-O-benzylidene-2-deoxy-2-phthalimido-β-D-glucopyranoside with ethyl 3-O-allyl-2-azido-4-O-benzyl-2,6-dideoxy-1-thio-α/β-l-glucopyranoside, under the agency of NIS/TfOH afforded methyl 3-O-(3-O-allyl-2-azido-4-O-benzyl-2,6-dideoxy-α-l-glucopyranosyl)-4,6-O-benzyli-dene-2-deoxy-2-phthalimido-β-D-glucopyranoside. Removal of the allyl function of the latter dimer, followed by condensation with properly protected 2-azido-2-deoxy-glucosyl donors, in the presence of suitable promoters, yielded selectively methyl 3-O-(3-O-[6-O-acetyl-2-azido-3,4-di-O-benzyl-2-deoxy-α-D-glucopyranosyl]-2-azido-4-O-benzyl-2,6-dideoxy-α-l-glucopyranosyl)-4,6-O-benzylidene-2-deoxy-2-phthalimido-β-D-glucopyranoside. Deacetylation and subsequent glycosylation of the free HO-6 with phenyl 2,3,4,6-tetra-O-benzoyl-1-seleno-β-D-glucopyranoside in the presence of NIS/TfOH furnished a fully protected tetrasaccharide. Deprotection then gave methyl 3-O-(3-O-[6-O-{β-D-glucopyranosyl}-2-acetamido-2-deoxy-β-D-glucopyranosyl)-2-acetamido-2,6-dideoxy-α-L-glucopyranosyl)-2-acetamido-2-deoxy-β-D-glucopyranoside.  相似文献   

3.
The known 1,3,4-oxathiazol-2-ones with crystal structures reported in the Cambridge Structural Database are limited (13 to date) and this article expands the library to 15. In addition, convenient starting materials for the future exploration of 1,3,4-oxathiazol-2-ones are detailed. An unexpected halogenated propanamide has also been identified as a by-product of one reaction, presumably reacting with HCl generated in situ. The space group of 5-[(E)-2-chloroethenyl]-1,3,4-oxathiazol-2-one, C4H2ClNO2S, ( 1 ), is P21, with a high Z′ value of 6; the space group of rac-2,3-dibromo-3-chloropropanamide, C3H4Br2ClNO, ( 2 ), is P21, with Z′ = 4; and the structure of rac-5-(1,2-dibromo-2-phenylethyl)-1,3,4-oxathiazol-2-one, C10H7Br2NO2S, ( 3 ), crystallizes in the space group Pca21, with Z′ = 1. Both of the structures of compounds 2 and 3 are modeled with two-component disorder and each molecular site hosts both of the enantiomers of the racemic pairs (S,S)/(R,R) and (R,S)/(S,R), respectively.  相似文献   

4.
Nitro derivatives of 1-R-1,2-benzoisothiazol-3-one 1-oxide were synthesized by the reactions of 2-alkyl(phenyl)thio-4-nitro- and 4,6-dinitro-2-(phenylthio)benzamides with chlorine in 60% acetic acid. Analogous reactions of 2-(n-butylthio)-4-nitro- and 2-(tert-butylthio)-4-nitrobenzamides with chlorine afforded 2-butyl- and 2-H-1,2-benzoisothiazol-3-one 1-oxides, respectively. The proposed reaction mechanism includes the formation and subsequent transformations of S-alkyl-S-aryl- and S,S-diarylchlorosulfonium chlorides.  相似文献   

5.
Cationic copolymerizations of 4-methyl-2-methylene-1,3-dioxane, 2 (M1), with 2-methylene-1,3-dioxane, 1 (M2); of 4,4,6-trimethyl-2-methylene-1,3-dioxane, 3 (M1), with 2-methylene-1,3-dioxane, 1 (M2); of 4-methyl-2-methylene-1,3-dioxolane, 5 (M1), with 2-methylene-1,3-dioxolane, 4 (M2); and of 4,5-dimethyl-2-methylene-1,3-dioxolane, 6 (M1), with 2-methylene-1,3-dioxolane, 4 (M2) were conducted. The reactivity ratios for these four types of copolymerizations were r1 = 1.73 and r2 = 0.846; r1 = 2.26 and r2 = 0.310; r1 = 1.28 and r2 = 0.825; r1 = 2.23 and r2 = 0.515, respectively. The relative reactivities of these monomers towards cationic polymerization are: 3 > 2 > 1; and 6 > 5 > 4. With both five- and six-membered ring cyclic ketene acetals, the reactivity increased with increasing methyl substitution on the ring. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 861–871, 1998  相似文献   

6.
Reactions of cyanomethanesulfonamides with aromatic aldehydes in the presence of AcOH and piperidine produced the addition products, the 1-cyano-2-arylethenesulfonamides, whereas reactions with benzonitrile yielded the 2-amino-1-cyano-2-phenylethenesulfonamides only when done in THF with BuLi. No addition products were isolated from the analogue reactions with 2-hydroxybenzaldehyde (salicylaldehyde). Instead, we obtained 2-imino-2H-chromene-3-sulfonamides with good to excellent yields. These 2H-chromene derivatives allowed a number of transformations, from which the reactions with orthoformates opened an approach to the hitherto unknown benzopyrano[3,2-e] [1,2,4]thiadiazine ring system.  相似文献   

7.
Stereochemistry of [2+2]- and [1+1]-macrocylization of cis-4-cyclohexene-1,2-dicarboxylic acid with 1,2-dibromoethane, 1,3-dichloro-2-propanol, and 1,5-dichloro-3-oxapentane was studied. The reaction of cis-4-cyclohexene-1,2-dicarboxylic acid with 1,2-dibromoethane according to the [2+2]-cyclization scheme gave a mixture of stereoisomeric macroheterocycles with cis,syn,cis- and cis,anti,cis-junction of the polyether and cyclohexene rings. In the reaction of cis-4-cyclohexene-1,2-dicarboxylic acid with 1,5-dichloro-3-oxapentane, a mixture of crown compounds with cis,anti,cis- and cis-junction of the polyether and cyclohexene rings was obtained. The structure of the products was established on the basis of their chemical transformations and spectral data.  相似文献   

8.
2-(9H-Xanthen-9-ylmethyl)-1H-benzimidazole ( 2a ) was prepared by condensing 9H-xanthene-9-acetic acid ( 1a ) with 1,2-benzenediamine. Similarly, 2-(9H-thioxanthen-9-ylmethyl)-1H-benzimidazole ( 2b ) and its S,S-dioxide ( 2d ) were obtained. Compound 2d was also prepared by oxidizing 2b with hydrogen peroxide in acetic acid. Heating of 9H-thioxanthene-9-acetic acid 10-oxide ( 1c ) with 1,2-benzenediamine gave 9-methylene-9H-thioxanthene ( 3 ). 2-(9H-Thioxanthen-9-ylmethyl)-1H-benzimidazole S-oxide ( 2c ) was obtained by oxidizing 2b with m-chloroperbenzoic acid in acetone.  相似文献   

9.
On reacting the 3-aminopyridazines 1a,d,e with dimethyl acetylenedicarboxylate (DMAD), the pyrimido[1,2-b]pyridazin-2-(2H)-ones 2e-g , whereas starting from 1f , the 4(4H)-ones 5a and 3b,d were prepared. In the 2(2H)-one series, the reactions of 2b with various amino compounds resulted in various types of products. The reaction of N-methylaminopyridazines 1g,h with DMAD led to the endo-N-substituted derivatives 8a,b , whereas 1h with diethyl ethoxymethylenemalonate (DEM) gave the exo-N-substituted compound 1k. The constitution of the compounds was proved by spectroscopic and chemical evidences.  相似文献   

10.
1,4-Diphenyl-2,3-dioxabicyclo[2.2.1]hept-5-ene ( 2 ), on treatment with a catalytic amount of trimethylsilyl trifluoromethanesulfonate (Me3SiOTf) in CH2Cl2 at ?78°, reacts with excess (?)-menthone ( 10 ) to give (1S,2S,4′aS,5R,7′aS)-4′a,7′a-dihydro-2-isopropyl-5-methyl-6′,7′-diphenylspiro[cyclohexane-1,3′-[7′H]cyclopenta-[1,2,4]trioxine] ( 11 ) and its (1R,2S,4′aR,5R,7′aR)-diastereoisomer 12 in a 1:1 ratio and in 21% yield. Repeating the reaction with 1.1 equiv. of Me3SiOTf with respect to 2 affords 11 , 12 , and (1S,2S,3′a.R,5R,6′aS)-3′a,6′a-dihydro-2-isopropyl-5-methyl-3′a-phenoxy-5′-phenylspiro[cyclohexane-l,2′-[4′H]cyclopenta[1,3]dioxole] ( 13 ) together with its(1R,2S,3′aS,5R,6′aR)-diastereoisomer 14 in a ratio of 3:3:3:1 and in 56% yield. (+)-Nopinone( 15 ) in excess reacts with 2 in the presence of 1.1 equiv. of Me3SiOTf to give a pair of 1,2,4-trioxanes ( 16 and 17 ) analogous to 11 and 12 , and a pair of 1,3-dioxolanes ( 18 and 19 ) analogous to 13 and 14 , in a ratio of 8:2:3:3 and in 85% yield. (?)-Carvone and racemic 2-(tert-butyl)cyclohexanone under the same conditions behave like 15 and deliver pairs of diastereoisomeric trioxanes and dioxolanes. In general, catalytic amounts of Me3SiOTf give rise to trioxanes, whereas 1.5 equiv. overwhelmingly engender dioxolanes. Adamantan-2-one combines with 2 giving only (4′aRS,7′aRS)-4′a,7′a-dihydro-6′.7′a-diphenylspiro[adamantane-2,3′-[7′H]cyclopenta[1,2,4]trioxine] in 98% yield regardless of the amount of Me3SiOTf used. The reaction of 1,4-dipheny 1-2,3-dioxabicyclo[2.2.2]oct-5-ene ( 32 ) with 10 and 1.1 equiv. of Me3SiOTf produces only the pair of trioxanes 33 and 34 homologous to 11 and 12 . Treatment of the (S,S)-diastereoisomer 33 with Zn and AcOH furnishes (1S,2S)-1,4-diphenylcyclohex-3-ene-1,2-diol. The crystal structures of 11 – 13 and 16 are obtained by X-ray analysis. The reaction courses of 10 and the other chiral cyclohexanones with prochiral endoperoxides 2 and 32 to give trioxanes are rationalized in terms of the respective enantiomeric silylperoxy cations which are completely differentiated by the si and re faces of the ketone function. The origin of the 1,3-dioxolanes is ascribed to 1,2 rearrangement of the corresponding trioxanes, which occurs with retention of configuration of the angular substituent.  相似文献   

11.
3-Methyl-3H-pyrazolo[1,5-d]tetrazoles 2 and 3-methyl-6-phenyl-3H-1,2,4-triazolo[1,5-d]tetrazole (4) have been unequivocally synthesized by annulation of the tetrazole moiety to the pyrazole resp. 1,2,4-triazole system. The constitution of some N-methyl substituted azolotetrazoles, formerly described as 3-methyl-3H-pyrazolo[1,5-d]tetrazoles 2, 3-methyl-6-phenyl-3H-1,2,4-triazolo[1,5-d]tetrazole (4) and 1-methyl-6-phenyl-1H-1,2,4-triazolo[4,3-d]tetrazole (5), has to be revised in favour of the corresponding mesoionic 2-methyl derivatives 2′, 4′, 5′. The structures of 3-methyl-3H- as well as of 2-methyl-2H-pyrazolo[1,5-d]tetrazole derivatives 2a, 2c, 2′a have been determined by X-ray analyses. The azapentalenic system is aromatic in all three measured compounds and mesoionic in the case of the 2-methyl-2H- substitution pattern. The phenyl and ester substituents are coplanar with the azapentalene system. 3-, 2-, and 1-Methylpyrazolo[1,5-d]tetrazoles exhibit different behaviour when allowed to react with stannous chloride or sodium ethoxide. Azolotetrazoles with a methyl substituent at N-1, N-2 or N-3 of the tetrazole moiety can be distinguished by a combination of 1H and 13C nmr with respect to the chemical shifts of the N-methyl group and the bridgehead carbon. Results of semiempirical calculations of the pyrazolo[1,5-d]tetrazole anion and of its N-methyl derivatives are discussed.  相似文献   

12.
The reaction of 3‐ethoxycarbonylmethylene‐3,4‐dihydroquinoxalin‐2(1H)‐one 5 with the Vilsmeier reagent, the treatment of 3‐(3,4‐dihydroquinoxalin‐2(1H)‐on‐3‐yl)‐1,2‐dihydro‐1,5‐benzodiazepin‐2(1H)‐one hydrochloride 7 with 10% sodium hydroxide and 3‐benzimidazoylquinoxaline‐2(1H)‐one 3 with both 1,2‐phenylenediamine dihydrochloride, and the reactions of 1,2‐phenylenediamine have been reinvestigated, and the structures of these reaction products have been revised. The aforementioned reactions have been shown to proceed with the formation of 1‐N,N‐dimethylaminomethylene‐2‐oxo‐1,2‐dihydrofuro[2,3‐b]quinoxaline 10 in the first case, the formation of 3‐[2‐(benzimidazol‐2‐on‐1‐yl)vinyl]‐1H‐quinoxalin‐2‐one 12 in the second case, and the formation of 2,3‐bis‐(1H‐benzimidazol‐2‐yl)quinoxaline 17 in the third case and not the formation of 3‐(N,N‐dimethylaminocarbonyl)furo[2,3‐b]quinoxaline hydrochloride 6 , the free base of 3‐(3,4‐dihydroquinoxalin‐2(1H)‐on‐3‐yl)‐1,2‐dihydro‐1,5‐benzodiazepin‐2(1H)‐one 7 , that is, compound 11 and benzodiazepine derivative 4 , as has been described earlier. In the third case, the formation of 2,3‐bis‐(1H‐benzimidazol‐2‐yl)quinoxaline 17 occurs according to the novel quinoxalin‐2(1H)‐one benzimidazole rearrangement discovered by us. The potential mechanisms for the investigated reactions are discussed.  相似文献   

13.
Noncovalent interactions of cis‐ and trans‐2‐butene, as the smallest model systems of molecules with cis and trans double bonds, were studied to find potential differences in interactions of these molecules. The study was performed using quantum chemical methods including very accurate CCSD(T)/CBS method. We studied parallel and displaced parallel interactions in 2‐butene dimers, in butane dimers, and between 2‐butene and saturated butane. The results show the trend that interactions of 2‐butene with butane are the strongest, followed by interactions in butane dimers, whereas the interaction in 2‐butene dimers are the weakest. The strongest calculated interaction energy is between trans‐2‐butene and butane, with a CCSD(T)/CBS energy of ?2.80 kcal mol?1. Interactions in cis‐2‐butene dimers are stronger than interactions in trans‐2‐butene dimers. Interestingly, some of the interactions involving 2‐butene are as strong as interactions in a benzene dimer. These insights into interactions of cis‐ and trans‐2‐butene can improve understanding of the properties and processes that involve molecules with cis and trans double bonds, such as fatty acids and polymers.  相似文献   

14.
The calculation of the 13C and 15N NMR chemical shifts by a combined molecular mechanics (Pcmodel 9.1/MMFF94) and ab initio (GIAO (B3LYP/DFT, 6-31 + G(d)) procedure is used to investigate the conformations of a variety of alkyl substituted anilines. The 13C shifts are obtained from the GIAO isotropic shielding (Ciso) with separate references for sp3 and sp2 carbons (δc = δref − Ciso). The 15N shifts are obtained similarly from the GIAO isotropic shielding (Niso) with reference to the 15N chemical shift of aniline. Comparison of the observed and calculated shifts provides information on the molecular conformations. Aniline and the 2,6-dialkylanilines exist with a rapidly inverting symmetric pyramidal nitrogen atom. The 2-alkylanilines have similar conformations with the NH2 group tilted away from the 2-alkyl substituent. The N,N-dialkylanilines show more varied conformations. N,N-dimethylaniline has a similar structure to aniline, but N-ethyl, N-methylaniline, N,N-diethylaniline, and N,N-diisopropylaniline are conformationally mobile with two rapidly interconverting conformers. In contrast, the anilines substituted at C2 and the nitrogen atom exist as one conformer where the steric interaction between the C2 substituent and the N substituent determines the conformation. In 2-methyl-N-methylaniline, the nitrogen atom is pyramidal as usual with the N-methyl opposite to the 2-methyl, but in 2-methyl-N,N-dimethyl aniline, the NMe2 group is now almost orthogonal to the phenyl plane. This is also the case with 2-methyl-N,N-diethylaniline and 2,6-diisopropyl-N,N-dimethylaniline. The comparison of the observed and calculated 15N chemical shifts confirms the above findings, in particular the pyramidal conformation of aniline and the above observations with respect to the conformations of the N,N-dialkylanilines.  相似文献   

15.
The preparation of two types of imidazole derivatives bearing a hydrazide group was achieved by treatment of the corresponding esters with NH2NH2?H2O in MeOH at room temperature. In the case of 4‐(ethoxycarbonyl)‐1H‐imidazole 3‐oxides 3 , hydrazides of type 1 were formed with retention of the N‐oxide structure (Scheme 1). Interestingly, due to a strong H‐bonding, no deoxygenation of the N→O function could be achieved even by treatment of 3 with Raney‐Ni. The second type, 2‐[(1H‐imidazol‐2‐yl)sulfanyl]acetohydrazides 2 , was obtained from 1H‐imidazole‐2(3H)‐thiones 4 in two steps via S‐alkylation with methyl bromoacetate, followed by treatment with NH2NH2?H2O (Scheme 2). An imidazole 7 , containing both types of hydrazide groups, was prepared analogously from ethyl 2,3‐dihydro‐2‐thioxo‐1H‐imidazole‐4‐carboxylate 4d (Scheme 4). Both types of hydrazides, 1 and 2 , were transformed successfully to the corresponding acylhydrazones 8 and 9 , respectively (Scheme 5). Furthermore, it has been shown that hydrazides of type 1 are useful starting materials for the synthesis of 1,2,4‐triazole‐3‐thiones 11 and 1,3,4‐thiadiazole‐2‐amines 12 , bearing an imidazole 3‐oxide moiety (Scheme 7).  相似文献   

16.
Hydrogen peroxide (H2O2) is popularly employed as a reaction reagent in cleaning processes for the chemical industry and semiconductor plants. By using differential scanning calorimetry (DSC) and vent sizing package 2 (VSP2), this study focused on the thermal decomposition reaction of H2O2 mixed with sulfuric acid (H2SO4) with low (0.1, 0.5 and 1.0 N), and high concentrations of 96 mass%, respectively. Thermokinetic data, such as exothermic onset temperature (T 0), heat of decomposition (ΔH d), pressure rise rate (dP/dt), and self-heating rate (dT/dt), were obtained and assessed by the DSC and VSP2 experiments. From the thermal decomposition reaction on various concentrations of H2SO4, the experimental data of T 0, ΔH, dP/dt, and dT/dt were obtained. Comparisons of the reactivity for H2O2 and H2O2 mixed with H2SO4 (lower and higher concentrations) were evaluated to corroborate the decomposition reaction in these systems.  相似文献   

17.
The reaction of aldimines with α‐(hydroxyimino) ketones of type 10 (1,2‐diketone monooximes) was used to prepare 2‐unsubstituted imidazole 3‐oxides 11 bearing an alkanol chain at N(1) (Scheme 2, Table 1). These products were transformed into the corresponding 2H‐imidazol‐2‐ones 13 and 2H‐imidazole‐2‐thiones 14 by treatment with Ac2O and 2,2,4,4‐tetramethylcyclobutane‐1,3‐dithione, respectively (Scheme 3). The three‐component reaction of 10 , formaldehyde, and an alkane‐1,ω‐diamine 15 gave the bis[1H‐imidazole 3‐oxides] 16 (Scheme 4, Table 2). With Ac2O, 2,2,4,4‐tetramethylcyclobutane‐1,3‐dithione or Raney‐Ni, the latter reacted to give the corresponding bis[2H‐imidazol‐2‐ones] 19 and 20 , bis[2H‐imidazol‐2‐thione] 21 , and bis[imidazole] 22 , respectively (Schemes 5 and 6). The structures of 11a and 16b were established by X‐ray crystallography.  相似文献   

18.
Synthetic routes leading to the preparation of 4-substituted 1,4-benzodiazepine-3,5-diones are described. Thus, 2-carbobenzoxyaminobenzoic acid was converted to its p-nitrobenzyl ester (I) and the decarbobenzoxylated product (II) gave, with ethyl α-bromoacetate, N-(2-carboxy p-nitrobenzylate)phenylglycine ethyl ester (III). The latter was hydrogenolyzed to N-(2-car-boxy)phenylglycine ethyl ester (IV), which was coupled with benzylamine to give N-(2-carboxy-benzylamido)phenylglycine ethyl ester (VIa). Saponification of VIa afforded N-(2-carboxy-benzylamido)phenylglycine (VIIa) which was cyclized with DCCI to produce 4-benzyl-2H-1,4-benzodiazepine-3,5(lH,4H)dione (VIIIa). Alternatively, 2-nitro-N-phenylbenzamide (Xb) was reduced to 2-amino-N-phenylbenzamide (XIb) which was converted to N-(2-carboxanih'do)-phenylglycine ethyl ester (VIb). The latter was converted to 4-phenyl-2H-1,4-benzodiazepine-3,5(1H,4H)dione (VIIIb) in an analogous fashion described for VIIIa.  相似文献   

19.
Triply excited states of many-electron atomic systems are characterized by the presence of strong electron correlation, closeness to more than one threshold, and degeneracy with many continua; therefore, they offer unusual challenges to theoretical methodologies. In the present article, we computed with reasonable accuracy all the n=2 intrashell triply excited states (2s22p 2P; 2s2p2 2D, 4P, 2P, 2S; and 2p3 2D, 2P, 4S) of three-electron atomic systems (Z=2, 3, 4, 6, 8, 10) by using a density functional formalism developed recently in our laboratory, based on the nonvariational Harbola–Sahni exchange potential in conjunction with a parametrized local Wigner and Lee–Yang–Parr correlation potentials. Nonrelativistic energies and densities are obtained by solving a Kohn–Sham-type differential equation. The calculated results are compared with available experimental and other theoretical data. The 2p3(4S)→1s2p2(4P) transition wavelength for the isoelectronic series is also computed. The overall good agreement of our results with the literature data indicates the reliability of the present density functional methodology for excited states of many-electron systems. © 1997 John Wiley & Sons, Inc. Int J Quant Chem 65 : 317–332, 1997  相似文献   

20.
Some condensation reactions of salicylaldehyde with various conjugated olefins, 1a, 1b, 1c, 2a, 2b, 2c , and 3 , were studied. In the condensations with 1a, 1b , and 1c gave 2,2-dimethyl-2H-chromene derivatives via “3–2 cyclization”, while the condensations with 2a, 2b, 2c , and 3 gave 2-methyl-2H-chromen-2-yl)acetic acid derivatives via “3–4 cyclization”.  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