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 50 毫秒
1.
Preparation of bis-heterazolidines bonded by a CH2, CH2–S–CH2 or CH2SCH2SCH2 groups through their nitrogen atoms is reported: 3-(1,3-oxazolidin-3-ylmethyl)-1,3-oxazolidine 1, 3-(4,4-dimethyl-1,3-oxazolidin-3-ylmethyl)-1,3-oxazolidine 2, 3-(1,3-diazolidin-3-ylmethyl)-1,3-diazolidine 3, 3-(1,3-thiazolidin-3-ylmethyl)-1,3-thiazolidine 4, 3-(1,3-thiazolidin-3-ylmethylsulfanylmethyl)-1,3-thiazolidine 5 and 3-(1,3-oxazolidin-3-ylmethylsulfanylmethyl-sulfanylmethyl)-1,3-oxazolidine 6. The solid state structures of 4 and 5 were determined by X-ray diffraction analyses. BH3–THF reduction reactions of compounds 1–6 were investigated. N→BH3 mono- and di-adducts of 1–6 were prepared and their structures calculated (ab initio 3-21G*).  相似文献   

2.
Reactions of acetal and 1,3-oxazolidine rings were examined using two kinds of iodosilane equivalent reagents, a 1:2 mixture of Me3SiNEt2 and MeI (reagent 1a) and a 1:1 mixture of Et3SiH and MeI containing a catalytic amount of PdCl2 (reagent 1b). In the reactions of alkanone ethylene acetals with reagent 1a, a C-O bond in the acetal ring readily cleaved to give 2-(trimethylsiloxy)ethyl enol ethers. Similarly, the C-O bond of 1,3-oxazolidine rings cleaved to give ring-opened imine or enamine derivatives. The reactions of aromatic ketone ethylene acetals and cyclohexanone trimethylene acetal led to deprotection of the acetal unit to liberate free ketones. With reagent 1b, cycloalkanone ethylene acetal afforded a dimeric product with 2-iodoethyl alkenoate moieties, while aromatic ketone ethylene or trimethylene acetals produced deprotected ketones.  相似文献   

3.
Nitric oxide synthase (NOS) has been divided into two major sub-enzymes, i.e. inducible NOS (iNOS) and constitutive NOS (cNOS). Although nitric oxide (NO) plays an important role as host defense mediator, excessive production of NO by iNOS has been involved in the pathology of many inflammatory diseases. Recently, we reported that the 2-imino-1,3-oxazolidine (1a) weakly inhibits iNOS and that introduction of an alkyl moiety on the oxazolidine ring of 1a enhances the inhibitory activity and selectivity for iNOS. In our search for better iNOS inhibitors, we focused our efforts on the 2-aminothiazole scaffold 3 as it possesses a ring similar to that of 1a. In this study, we evaluated the inhibitory activity of a series of 2-aminothiazole derivatives against both iNOS and neuronal NOS (nNOS). Our results show that introduction of appropriately-sized substituents at the 4- and 5-position of the 2-aminothiazole ring improves the inhibitory activity and selectivity for iNOS. We also found that the selectivity of 5a [5-(1-methyl)ethyl-4-methylthiazol-2-ylamine] and 5b [5-(1,1-dimethyl)ethyl-4-methylthiazol-2-ylamine] for iNOS was similar to that of oxazolidine derivative 1b (4-methyl-5-propyl-2-imino-1,3-oxazolidine) and much higher than that of L-NAME. However, we could not enhance the inhibitory activity against iNOS by introducing an alkyl substituent into the 2-aminothiazole ring as we could in the case of oxazolidine one. On the other hand, introduction of bulky or hydrophilic substituent at any position of the 2-aminothiazole ring remarkably decreased or even abolished the inhibitory activity against NOS.  相似文献   

4.
A substituted butadiene — 1-phenyl-4-(5-methyl-4-phenyl-2-pyridyl)buta-1,3-diene with a trans, trans configuration — was obtained by condensation of 2,5-dimethyl-4-phenylpyridine with cinnamaldehyde. Two 3-methyl-9-cinnamylidene-2-azafluorene isomers are formed as a result of condensation of the same aldehyde with 3-methyl-2-azafluorene. Data from the PMR and IR spectra were used to prove the configuration of the compounds obtained. It was established that the condensation of 3-methyl-2-azafluorene with salicylaldehyde gives 3-methyl-9-(2-hydroxybenzylidene)-2-azafluorene, which has a zwitterionic structure, and 1,2-bis(3-methyl-2-aza-9-fluorenylidene)ethane. Ideas regarding the chemical mechanism of the formation of the latter are presented. The preparation of an unsaturated alcohol — 3-methyl-9-allyl-2-aza-9-fluorenol — is described.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 1, pp. 82–86, January, 1978.  相似文献   

5.
The rapid parallel synthesis and characterization of diverse chirally defined 1,3-oxazolidines is reported. Three diversity elements were incorporated in a 6 x 4 x 4 block approach to generate a 96-member 1,3-oxazolidine library. The synthetic route involved initial attachment of six nonracemic phenylglycidols, (2S,3S)1A-C and (2R,3R)-2A-C, to 2% cross-linked polystyrene resin via a chlorodiethylsilane linker (PS-DES), followed by regio- and stereoselective oxirane ring opening with four primary amines (3a-d). The key condensation reaction between the resulting polymer-bound beta-amino alcohols and four aldehydes (4a-d) was found to occur optimally in warm benzene (60 degrees C) in the presence of anhydrous magnesium sulfate. Cleavage of the oxazolidines from the resin support was achieved with TBAF to give the individual members (2R,4R,5R)-5Aaa-Cdd and (2S,4S,5S)-6Aaa-Cdd in good to excellent yields (51-99%) based on mass recovery. Purities of all these crude products was generally >85% (as measured by LCMS). 1H, 13C NMR, and 1D difference nOe of the library members confirmed the structural and stereochemical integrity of the substituents around the 1,3-oxazolidine core. The asymmetric induction at C-2 (cis or trans to the C-4 substituent) ratio ranged from 4 to I to 49 to 1 across the library. This report highlights the versatility of the 1,3-oxazolidine heterocycle as a scaffold for concise parallel library construction and opens the way for high-throughput screening of such compounds in the biological sphere.  相似文献   

6.
The reaction of 5-methyl-5-(4-methyl-3-pentenyl)-4-methylene-1,3-dioxolan-2-one with primary amines gives the corresponding 4-hydroxy-4-methyloxazolidin-2-ones. These undergo an intramolecular amidoalkylation reaction to form 1-oxa-3-azapentalen-2-ones which are potential precursors of cyclopentanyl cis-2-amino alcohols.  相似文献   

7.
The compounds 2-(ethoxycarbonylmethyl)-2-l,3-oxazoline and 1-methyl-2-(methoxycarbonylmethylene)-1,3-oxazolidine are hydrochlorinated at the C=N or C= C bond, respectively, with subsequent opening of the oxazolidine ring. Depending on the reaction conditions, this occurs regioselectively to form various products.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 5, pp. 684–686, May, 1992  相似文献   

8.
The reaction of 4,5-dimethyl-2,2,2-trimethoxy-2,2-dihydro-1,3,2-dioxaphospholene with carbomethoxy, carbopropoxy and N,N-diphenylcarbamyl isocyanates yields, respectively, 2-methoxy-, 2-propoxy- and 2- diphenylamino-5-acetyl-5-methyl-2-oxazolin-4-one. The reaction with carbophenoxy isocyanate gives two products in a proportion which depends on experimental conditions: 2-phenoxy-5-acetyl-5-methyl-2-oxalin-4-one (1:1 stoichiometry) and 1,3-dicarbophenoxy-5-acetyl-5-methyl-hydantoin (1:2 stoichiometry). The 2-substituted 4- oxazolones are hydrolyzed to 5-acetyl-5-methyl-oxazolidin-2,4-dione. The alkyl group of the 2-alkoxy-4-oxazolones migrates to the adjacent nitrogen to give 3-alkyl-5-acetyl-5-methyl-oxazolidin-2,4-diones. The dioxaphospholene reacts with 2-substituted 2-thiazolin-4,5-diones to give 2-substituted 5-acetyl-5-methyl-2-thiazolin-4-ones, including rhodanine derivatives.  相似文献   

9.
Methyl- and phenylhydrazines react with 2-(diphenylacetyl)-1,3-indandione ( 1 ) to yield respectively the 1-(methylhydrazone) and the 1-(phenylhydrazone) of 2-(diphenylacetyl)-1,3-indandione ( 2a and 2b ). In comparison, acetic and benzoic acid hydrazides react with 1 to give respectively the α-(acetylhydrazone) and the α-(benzoylhydrazone) of 2-(diphenylacetyl)-1,3-indandione ( 3a and 3b ). Cyclization of 2a and 2b gives 2,3-disubstituted indeno[1,2-c]pyrazol-4(2H)-ones ( 7a and 7b ). Cyclization of 3a and 3b , followed by methylation, gives 1-methyl- and 2-methyl-3-(diphenylmethyl)indeno[1,2-c]pyrazol-4(1 and 2H)-ones ( 9a and 9b ). 2-Isovaleryl-1,3-indandione reacts with phenylhydrazine to give directly 3-isobutyl-1-phenylindeno[1,2-c]-pyrazol-4(1H)-one ( 10 ).  相似文献   

10.
Zwitterionic titanoxanes {Cp[η5-C5H4B(C6F5)3]Ti}2O (I) and {(η5-iPrC5H4)[η5-1,3-iPrC5H3B(C6F5)3]Ti}2O (II), which contain two positively charged Ti(IV) centres in the molecule, are able to catalyse the ring-opening polymerization of -caprolactone (-CL) in toluene solution and in bulk. The process proceeds with a noticeable rate even at room temperature and accelerates strongly on raising the temperature to 60 °C. The best results have been obtained on carrying out the reaction in bulk. Under these conditions, the use of I as a catalyst (-CL:I = 1000:1) gives at 60 °C close to quantitative yield of poly--CL with the molecular mass of 197 000. An increase in the -CL:I ratio to 6000:1 increases the molecular mass of poly--CL to 530 000. Tetrahydrofuran (THF) is also polymerized under the action of I albeit with a lesser rate. However, the molecular mass of the resulting poly-THF can reach rather big values under optimal conditions (up to 217 000 at 20 °C and the THF:I ratio of 770:1). A rise in the reaction temperature from 20 to 60 °C results here to a decrease in the efficiency of the process. Titanoxane II is close to I in its catalytic activity in the -CL polymerization but it is much less active in the polymerization of THF. Propylene oxide (PO), in contrast to -CL and THF, gives with I only liquid oligomers in wide temperature and PO:I molar ratio ranges (−30 to +20 °C, PO:I = 500–2000:1). γ-Butyrolactone and 1-methyl-2-pyrrolidone are not polymerized under the action of I at room temperature. The reactions found are the first examples of catalysis of the cationic ring-opening polymerization by zwitterionic metallocenes of the group IVB metals.  相似文献   

11.
An investigation by 220-MHz NMR spectroscopy was carried out on the alternating copolymers of acrylic monomer with 2-substituted 1,3-diolefin. The chain structures were determined. The acrylic monomers used were methyl methacrylate (MMA), acrylonitrile (AN), and methacrylonitrile (MAN); isoprene (IP) and chloroprene (CLP) were the 1,3-diolefins. In the MAN–IP alternating copolymer, the 1-position methylene protons of IP showed an AB quartet peak, confirming the α-1 linkage structure. Similarly, in the MMA–CLP and AN–CLP copolymers, the 1-position methylene protons of CLP showed and AB quartet and an ABX pattern, respectively, confirming the α-1 linkage structure in both these cases also. The α-1 linkage structure was also revealed by the decoupling technique in the MAN–CLP alternating copolymer. The AN–IP and MMA–IP alternating copolymers also possess a bond between the α-position of the acrylic monomer and the 1-position of IP. The monomeric units in the alternating copolymers of acrylic monomers with 2-substituted 1,3-diolefins were generally linked at the α-position of acrylic monomer and the 1-position of 1,3-diolefin. On the other hand, in the Diels-Alder adducts of acrylic monomer with 2-substituted 1,3-diolefin, the reaction takes place between the α-position of acrylic monomer and the 4-position of 1,3-diolefin. The regioselectivity of the alternating copolymers and the Diels-Alder adducts is quite compatible with the expectations from molecular orbital theory.  相似文献   

12.
A fast atom bombardment (FAB) mass spectrometric study on the open-chain compound 1,3-bis(cyanomethylenetriphenylphosphorane)propane-1,3-dione and on the cyclic zwitterionic compounds 4-oxy-5-triphenylphosphonium-6-methyl-2-pyrone and 4-oxy-5-triphenylphosphonium-6-phenyl-2-pyrone, obtained by reaction of carbon suboxide, C3O2, with stabilized phosphorus ylides, Ph3PCHX (X?CN, COMe, COPh), is described. The FAB mass spectrometric behaviour of these compounds is compared with that shown by tri-phenylphosphoranilideneketene, Ph3P ? C ? C ? O, and by 4-hydroxy-6-methyl-2-pyrone, with the aid of metastable ion data and collision spectroscopy.  相似文献   

13.
甜菜碱型两性离子聚合物P(AM-DMAPAAS)的盐溶液性质   总被引:1,自引:0,他引:1  
丁伟  毛程  韦兆水  李明  于涛  曲广淼 《应用化学》2011,28(5):555-559
将丙烯酰胺丙基二甲基胺(DMAPAA)和1,3-丙基磺内酯,在55 ℃下反应20 h,合成了3-(丙烯酰胺丙基二甲胺基)丙磺酸盐(DMAPAAS),将其在盐溶液中与丙烯酰胺(AM)单体进行自由基共聚合反应,获得净电荷为零的磺基甜菜碱型两性离子共聚物P(AM-DMAPAAS);对该两性离子共聚物进行了表征和溶解性评价。 研究结果表明,共聚物在NaCl溶液中的粘度比在纯水中的大,在Mg2+和Ca2+盐溶液中的粘度更大,且随着溶液浓度的增大而增大,表现出明显的反聚电解质溶液性质。 升高相同温度,两性离子共聚物的粘度保留率是普通聚丙烯酰胺的1.4倍。  相似文献   

14.
The reaction of cotarnine with 1,3-dimethylbarbituric acid afforded 5-(8-methoxy-2-methyl-6,7-methylenedioxy-1,2,3,4-tetrahydroisoquinolin-1-yl)-1,3-dimethylbarbituric acid. Its crystal structure was established by X-ray diffraction analysis and the structure in solution was studied by 1H NMR spectroscopy. This compound has a zwitterionic structure. In the crystal, the molecules are linked in dimers by intermolecular hydrogen bonds, and in solutions, the dimers occur in equilibrium with the monomers.  相似文献   

15.
5,6-Unsubstituted 1,3-oxazine-2,4-diones ( 3 ) and 6-unsubstituted 5-methyl-1,3-oxazine-2,4-diones ( 4 ) were prepared by reduction of the corresponding 6-chloro derivatives ( 1 and 2 ). Treatment of 6-chloro-3-methyl-1,3-oxazine-2,4-dione ( 1a ) with sodium azide, sodium cyanide, secondary amines and aniline gave the corresponding 6-substituted compounds ( 7, 9, 10 and 11 ) while the reaction of 1a and 2a,b with primary aliphatic amines such as methylamine and ethylamine caused a ring transformation to pyrimidine ring system giving barbituric acids ( 13a-d ).  相似文献   

16.
2-Methoxy-6-ethoxy-5-cyano-3,4-dihydro-2H-pyran ( 1 a), 2-isobutoxy-6ethoxy-5-cyano-3,4-dihydro-2H-pyran ( 1 b), and 2,6-diethoxy-3-methyl-5-cyano-3,4-dihydro-2H-pyran ( 1 c) were prepared by (4 + 2) cycloaddition reaction of ethyl α-cyanoacrylate with the corresponding vinyl ethers. Compounds 1 a-c were ring-open polymerized by cationic catalyst to obtain alternating head-to-head (H? H) copolymers. For comparison, head-to-tail (H? T) copolymer 3 a was also prepared by free radical copolymerization of the mixture of the corresponding monomers. The H–H copolymer exhibited minor differences in its 1H-NMR and IR spectra, but in the 13C-NMR spectrum significant differences were shown between the H? H and H? T copolymers. Glass transition temperature (Tg) of H? H copolymer was higher than that of the corresponding H? T copolymer, but thermal decomposition temperature of the H? H copolymer was lower than that of the H? T copolymer. Compounds 1 a and 1 b copolymerized well with styrene by cationic catalyst, but compound 1 c failed to copolmerize with styrene. All of the H-H and H-T copolymers were soluble in common solvents and the inherent viscosities were in the range 0.2–0.5 g/dL.  相似文献   

17.
The methanesulfonic acid catalyzed reaction of 1-(4-chloro- and 2,4-dichlorophenyl)-2-(1-methyl-2-imida-zolyl)ethanones 1a and 1b with glycerol produced cis- and trans-{2-haloaryl-2-[(1-methyl-2-imidazolyl)methyl]-4-hydroxymethyl}-1,3-dioxolanes 2a and 2b with a 2:1 cis/trans ratio. Besides these five-membered ketals, the reaction of 1a with glycerol afforded a small amount of trans-{2-(4-chlorophenyl)-2-[(1-methyl-2-imidazolyl)methyl]-5-hydroxy}-1,3-dioxane ( 3a , 7%). The reaction of methanesulfonyl chloride with cis-1 formed the corresponding methanesulfonates, cis- 4 , which rapidly cyclized to the title compounds 5 . Base-catalyzed ring opening of 5 furnished 1-methyl-5,6-dihydro-6-hydroxymethyl-8-(4-chloro- and 2,4-dichlorophenyl)-1H-imidazo[3,2-d][1,4]oxazepinium methanesulfonates 7 . Acid-catalyzed hydrolyses of 5 or 7 provided 1-methyl-2-[(4-chloro- and 2,4-dichloro)phenacyl]-3-[(2,3-dihydroxy)-1-propyl]imidazolium salts 12 . Structure proofs were based on extensive 1H and 13C chemical shifts and coupling constants and structures of 3a and 5a were confirmed by single crystal X-ray crystallography.  相似文献   

18.
5-Benzoyl-4-methylpyrimidines 4a,b and 5-acetyl-4-phenylpyrimidines 5a,b reacted with hydrazines in alcoholic acidic medium to give respectively 4-acetyl-3-phenylpyrazoles 7, 9 and 10 and 4-benzoyl-3-methylpyrazoles 6, 8 and 11 . In the reaction with phenylhydrazine, 5-benzoyl-4-methyl-2-methylthiopyrimidine ( 4a ) led exclusively to 4-acetyl-1,3-diphenylpyrazole ( 10 ) as 5-acetyl4-phenyl-2-methylthiopyrimidine ( 5a ) led to 4-benzoyl-3-methyl-1-phenylpyrazole ( 11 ) via the initial formation of phenylhydrazones of pyrimidines 4a and 5a . However, 5-benzoyl-4-methyl-2-phenylpyrimidine ( 4b ) and 5-acetyl-2,4-diphenylpyrimidine ( 5b ) reacted with phenylhydrazine to afford, each of them, a mixture of two isomeric pyrazoles. The mechanism of these ring contraction reactions is discussed.  相似文献   

19.
The AlCl3-initiated cationic polymerization of optically active 1-olefins yields polymers of varying optical rotatory power. Polymers of (+)-3-methyl-1-pentene and (?)-4-methyl-1-hexene prepared between ?78 and ?55°C. in CH2Cl2 or n-heptane are almost completely optically inactive. Under identical reaction conditions (+)-5-methyl-1-heptene gives polymers of significant optical rotatory power. Alternating SO2copolymers of the same olefins, formed in reactions which proceed through free-radical intermediates, yield optically active products with specific rotations similar to those of low molecular weight analogs. These results are consistent with a cationic polymerization mechanism in which the growing chain undergoes intramolecular hydride shift and the asymmetric carbon atoms are converted into carbonium ions. The data also provide evidence for the lack of rearrangement in free-radical polymerization. By comparing the specific rotations of the cationic and free-radical polymers, the extent of rearrangement during cationic polymerization can be estimated. The calculations show that the 1,2-polymer in cationic poly-3-methyl-1-pentene is less than 2%, the sum of 1,2- and 1,3-polymer in cationic poly-4-methyl-1-hexene is less than 4%, and the sum of 1,2-, 1,3-, and 1,4-polymer in cationic poly-5-methyl-1-heptene is 14–20%.  相似文献   

20.
The synthesis and characterization of copolymers from styrene and 1,3‐pentadiene (two isomers) are reported. Styrene/1,3‐pentadiene (1:1) copolymerization with carbanion initiator yield living, well‐defined, alternating (r1 = 0.037, r2 = 0.056), and highly stereoregular copolymers with 90%–100% trans‐1,4 units, designed Mns and low ÐMs (1.07–1.17). The first‐order kinetic resolution and NMR spectra demonstrate that the copolymers obtained possess strictly alternating structure containing both 1,4‐ and 4,1‐enchaiments. Also a series of copolymers with varying degrees of alternation are synthesized from para‐alkyl substituted styrene derivatives and 1,3‐pentadiene. The degree of alternation is strongly dependent on the polarity of solvent, reaction temperature, type of transcis isomer of 1,3‐pentadiene and para‐substituted group in styrene. The macro zwitterion forms (SPC) through the distribution of electronic charges from the donor (1,3‐pentadiene) to the acceptor (styrenes) are proposed to interpret the carbanion alternating copolymerization mechanism. Owing to the versatility of the carbanion‐initiating reaction, the present alternating strategy based on 1,3‐pentadiene (especially cis isomer) can serve as a powerful tool for precise control of polymer chain microstructure, architecture, and functionalities in one‐pot polymerization.

  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