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 39 毫秒
1.
13(S)-hydroperoxy- and 13(S)-hydroxyoctadeca-9,11-dienoic acids (1a/b), 15(S)-hydroperoxy- and 15(S)-hydroxyeicosa-5,8,11,13-tetraenoic acids (2a/b), and their methyl esters reacted smoothly with NO2- in phosphate buffer at pH 3-5.5 and at 37 degrees C to afford mixtures of products. 1b methyl ester gave mainly the 9-nitro derivative 3b methyl ester (11% yield) and a peculiar breakdown product identified as the novel 4-nitro-2-oximinoalk-3-enal derivative 4 methyl ester (15% yield). By GC-MS hexanal was also detected among the products. Structures 3b and 4 methyl esters were secured by 15N NMR analysis of the products prepared from 1b methyl ester upon reaction with Na15NO2. 4 methyl ester (14% yield) was also obtained from 1a methyl ester along with the nitrated hydroperoxy derivative 3a methyl ester (10% yield). Under the same conditions, 2a/b methyl esters gave mainly the corresponding nitrated derivatives 5a/b, with no detectable breakdown products, whereas the model compound (E,E)-2,4-hexadienol (6) afforded two main nitrated derivatives identified as 7 and 8. A reaction pathway for 1a/b methyl esters was proposed involving conversion of nitronitrosooxyhydro(pero)xy intermediates which would partition between two competing routes, viz., loss of HNO2, to give 3a/b methyl esters, and a remarkably facile fission leading to 4 methyl ester and hexanal.  相似文献   

2.
Charge-transfer-induced decomposition (CTID) of bicyclic dioxetanes 1b-d bearing a 3-hydroxylphenyl moiety substituted with a benzothiazol-2-yl group at the 2-, 6-, or 5-position was investigated, and their chemiluminescence properties were compared to each other, based on those for a 4-benzothiazolyl analogue 1a. Dioxetanes 1c and 1d underwent CTID to give the corresponding oxido anions of keto esters 8c or 8d in the singlet excited state with high efficiencies similarly to the case of 1a. On the other hand, 1b showed chemiluminescence with quite low efficiency, though it gave exclusively keto ester 2b. The marked decline of chemiluminescence efficiency for 1b was attributed to 1b mainly being decomposed to 8b through a radiationless pathway, in which intramolecular nucleophilic attack of nitrogen in the benzothiazolyl group to dioxetane O-O took place to give cyclic intermediate cis-11.  相似文献   

3.
The new types of ferrocenyloxaaliphatic acid ester, FcCHROCHR′COOMe (R = H, Me, Ph; R′ = H, Me) (7) have been prepared by the action of alkoxides derived from methyl glycolate or methyl lactate on the corresponding ferrocenylcarbinyl acetates (2) or N,N,N-trimethylferrocylammonium iodides (4). The esters obtained were accompanied by a small quantity of oligomeric esters, FcCHR(OCHR′CO)n OMe (9), and with more or less ferrocyl methyl ethers (8). As opposed to the alkaline hydrolysis of the analogous methyl benzoxyacetate (6) into benzoxyacetic acid (5) the acidification of sodium alkanoates 10 obtained by saponification of esters 7 gave unexpectedly the corresponding ferrocenylcarbinols 1. In a similar way the esters 7 were converted into mixtures of the mentioned carbinols and diferrocyl ethers 11 under action of aqueous hydrochloric acid. The mechanisms of the reactions 10 → 1 and 7 → 1, 11 are discussed.  相似文献   

4.
Trimethylsilyl and Trimethylstannyl Esters of the Trithiophosphonic Acids; Preparation, Protolysis, and Further Reactions The organotrithiophosphonic acid bis(trimethylsilyl) esters RP(S)(SSiMe3)2, R = Ph ( 1a ), R = t-Bu ( 1b ) and R = Me ( 1c ) are formed in high yield from the organo-bis(trimethylsilyl)-phosphanes RP(SiMe3)2 by the addition of three sulfur atoms (3/8 S8) in toluene solution. 1b has also been prepared by reacting disodium tert-butyltrithiophosphonate, Na2t-BuP(S)S2 ( 2 ) with SiMe3Cl. Analogous reactions can be used for the preparation of the stannyl esters RP(S)(SSnMe3)2, R = t-Bu ( 5a ) and R = Me ( 5b ). More favorable, however, these compounds are synthesized from the corresponding silyl esters 1b, c and SnMe3Cl or, as has been shown in the case of 5a , by reacting dithiophosphonic acid anhydrides (RP(S)S)2 with (SnMe3)2S. Low temperature solvolysis of the silyl esters 1b, c with water or methanol results in the free organotrithiophosphonic acids RP(S)(SH)2, R = t-Bu ( 6a ) and R = Me ( 6b ), which can be isolated as white solids at —30°C. Contrary to the sodium salt 2 and the stannyl esters 5a, b , the acids 6a, b , and, to a smaller extent, the silyl esters 1a—c are thermally not stable towards decomposition into the thioanhydrides (RP(S)S)2. The silyl and stannyl esters 1a—c and 5a, b , respectively, are capable of cleaving the ether linkage of tetrahydrofuran, which in the case of 1b leads to the quantitative formation of t-BuP(S)[S(CH2)4OSiMe3]2 ( 9 ). 1b, c and 5a, b react with Cl2 and Br2 forming the 3,6-diorgano-3,6-dithio-1,2,4,5,3,6-tetrathiadiphosphorinanes (RP(S)S2)2, R = t-Bu ( 7a ) and R = Me ( 7b ).  相似文献   

5.
The Deltadelta (deltaS-deltaR) values for the C-1 methyl 1H signals in the 1H-NMR spectroscopy of the bis-MTPA esters of four synthetic stereoisomers of alkane-6,8-diols, viz., bis-MTPA esters of (6S,8R)-C27- (1a) and C29- (3a) (Deltadelta = -0.05 ppm), (6R,8S)-C27- (2a) and C29- (4a) (Deltadelta = +0.05 ppm), (6S,8S)-C27- (5a) (Deltadelta = -0.01 ppm), and (6R,8R)-C27- (6a) (Deltadelta = +0.01 ppm) alkane-6,8-diols, made it possible to differentiate unequivocally among the four stereoisomers. This allowed the determination of the (6S,8R)-stereochemistry (Deltadelta = -0.05 ppm for the bis-MTPA esters) for the natural C27- and C29-alkane-6,8-diols isolated from the flowers of three Compositae plants, Carthamus tinctorius, Cynara cardanclus, and Taraxacum platycarpum.  相似文献   

6.
A Diels-Alder cyclization proposed to occur during polyketide synthase assembly of the bicyclic core of lovastatin (1) (mevinolin) by Aspergillus terreus MF 4845 was examined via the synthesis of the N-acetylcysteamine (NAC) thioester of [2,11-(13)C(2)]-(E,E,E)-(R)-6-methyldodecatri-2,8,10-enoate (5a). In vitro Diels-Alder cyclization of the corresponding unlabeled NAC ester 5b, ethyl ester 18b, and acid 20b yielded two analogous diastereomers in each case, under either thermal or Lewis acid-catalyzed conditions. The reaction of thioester 5 proceeds readily at 22 degrees C in aqueous media. For 18b, one product is trans-fused ethyl (1R,2R,4aS, 6R,8aR)-1,2,4a,5,6,7,8,8a-octahydro-2,6-dimethylnaphthalene-1-carboxylate (30) (endo product), and the other is cis-fused ethyl (1R,2S,4aR,6R,8aR)-1,2,4a,5,6,7,8,8a-octahydro-2,6-dimethylnaphthalene-1-carboxylate (31) (exo product). Isomer 21 with stereochemistry analogous to 4a,5-dihydromonacolin L (2), a precursor of 1, was made by transformation of a tricyclic lactone, (1S,2S,4aR,6S,8S,8aS)-1-(ethoxycarbonyl)-1,2,4a,5,6,7,8,8a-octahydro-2-methyl-6,8-naphthalenecarbolactone (22) using reduction and Barton deoxygenation. Comparison of 21 with 30 and 31 confirmed the structural assignments and showed that the nonenzymatic 4 + 2 cyclizations of 5, 18, and 20 proceed via chairlike exo and endo transition states with the methyl substituent pseudoequatorial. The proposed biosynthetic Diels-Alder leading to lovastatin (1) would require an endo conformation with the methyl substituent pseudoaxial. Intact incorporation of the labeled hexaketide triene 5a into 1 was not achieved because of rapid degradation by A. terreus cells.  相似文献   

7.
In aqueous media alpha-keto amides 4-YC6H4OCH2COCON(R)CH(R')CH3 (5a, R = Et, R' = H; 5b, R = iPr, R' = Me) with para-substituted phenolic substituents (Y = CN, CF3, H) undergo photocleavage and release of 4-YC6H4OH with formation of 5-methyleneoxazolidin-4-ones 7a,b. For both 5a,b quantum yields range from 0.2 to 0.3. The proposed mechanism involves transfer of hydrogen from an N-alkyl group to the keto oxygen to produce zwitterionic intermediates 8a-c that eliminate the para-substituted phenolate leaving groups. The resultant imminium ions H2C=C(OH)CON+(R)=C(R')CH3 9a,b cyclize intramolecularly to give 7a,b. The quantum yields for photoelimination decrease in CH3CN, CH2Cl2, or C6H6 due to competing cyclization of 8a,b to give oxazolidin-4-one products which retain the leaving group 4-YC6H4O- (Y = H, CN). A greater tendency to undergo cyclization in nonaqueous media is observed for the N,N-diethyl amides 5a than the N,N-diisopropyl amides 5b. With para electron releasing groups Y = CH3 and OCH3 quantum yields for photoelimination significantly decrease and 1,3-photorearrangment of the phenolic group is observed. The 1,3-rearrangement involves excited state ArO-C bond homolysis to give para-substituted phenoxyl radicals, which can be observed directly in laser flash photolysis experiments.  相似文献   

8.
Since 2,6-dimethyl-4-aryl-1,4-dihydropyridine 3,5-diesters themselves are not hydrolyzed by commercially available hydrolases, derivatives with spacers containing a hydrolyzable group were prepared. Seven acyloxymethyl esters of 5-methyl- and 5-(2-propoxyethyl) 4-[2-(difluoromethoxy)phenyl]-2,6-dimethyl-1,4-dihydro-3,5-pyridinedicarboxylate were synthesized and subjected to Candida rugosa lipase (CRL) catalyzed hydrolysis in wet diisopropyl ether. A methyl ester at the 5-position and a long or branched acyl chain at C3 gave the highest enantiomeric ratio (E values). The most stereoselective reaction (E = 21) was obtained with 3-[(isobutyryloxy)methyl] 5-methyl 4-(2-difluoromethoxyphenyl)-2,6-dimethyl-1,4-dihydropyridine-3,5-dicarboxylate, and this compound was used to prepare both enantiomers of 3-methyl 5-(2-propoxyethyl) 4-[2-(difluoromethoxy)phenyl]-2,6-dimethyl-1,4-dihydro-3,5-pyridinedicarboxylate. The absolute configuration of the enzymatically produced carboxylic acid was established to be 4R by X-ray crystallographic analysis of its 1-(R)-phenylethyl amide.  相似文献   

9.
Selective addition of the chiral, sulfonimidoyl substituted bis(allyl)titanium complexes 5a-d, which are configurationally labile in regard to the Calpha-atoms, to N-toluenesulfonyl (Ts)-, N-2-trimethylsilylethanesulfonyl (SES)-, and N-tert-butylsulfonyl (Bus) alpha-imino ester (9a-c) in the presence of Ti(OiPr)(4) and ClTi(OiPr)(3) afforded with high regio- and diastereoselectivities in good yields the (syn, E)-configured beta-alkyl-gamma,delta-unsaturated alpha-amino acid derivatives 2a-g, which carry a chiral, electron-withdrawing nucleofuge at the delta-position and a cyclohexyl, an isopropyl, a phenyl, and a methyl group at the beta-position. Addition of the cyclic bis(allyl)titanium complex 14 to N-Bus alpha-imino ester 9c afforded with similar high regio- and diastereoselectivities the (E)- and (Z)-configured amino acid derivatives (E)-8 and (Z)-8. Reaction of complexes 5a-d with alpha-imino esters 9a-c in the presence of Ti(OiPr)(4) occurs stepwise to give first the mono(allyl)titanium complexes containing 2a-g as ligands, which react in the presence of ClTi(OiPr)(3) with a second molecule of 9a-c with formation of two molecules of 2a-g. Formation of (S,R,E)-configured homoallylic amines 2a-g entails Si,Re,E processes of alpha-imino esters 9a-c with the (R,R)-configured bis(allyl)titanium complexes (R,R)-5a-d and (R)-configured mono(allyl)titanium complexes (R)-17a-d, both of which are most likely in rapid equilibrium with their (S,S)-diastereomers and (S)-diastereomers, respectively. Interestingly, in the reaction of 5a-d with aldehydes, the (S,S)-configured complexes (S,S)-5a-d are the ones which react faster. Reaction of the N-titanated amino acid derivatives Ti-2a and Ti-2b with N-Ts alpha-imino ester 9a led to the highly diastereoselective formation of imidazolidinones 15a and 15b, respectively. Cleavage of the sulfonamide group of the N-Bus amino acid derivative 2d with CF(3)SO(3)H gave quantitatively the sulfonimidoyl functionalized amino acid H-2d. A Ni-catalyzed cross-coupling reaction of the amino acid derivative 2e with ZnPh(2) led to a substitution of the sulfonimidoyl group by a phenyl group and furnished the enantiomerically pure protected alpha-amino acid Bus-1. Two new N-sulfonyl alpha-imino esters, the SES and the Bus alpha-imino esters 9b and 9c, respectively, have been synthesized from the corresponding sulfonamides by the Kresze method in medium to good yields. The N-SES alpha-imino ester 9b and the N-Bus alpha-imino ester 9c should find many synthetic applications, in particular, in cases where the N-Ts alpha-imino ester 9a had been used before.  相似文献   

10.
5-氟尿嘧啶N1-甲酰基氨基酸、短肽的合成及抗肿瘤活性   总被引:4,自引:0,他引:4  
5-氟尿嘧啶N1-甲酰氯分别与Gly、Val、Leu、Ile、Phe、Asp和Glu的苄酯反应,制备了7种5-氟尿嘧啶-N1-甲酰基氨基酸苄酯。氢解后得到了相应的5-氟尿嘧啶N1-甲酰氨基酸。将其进一步与氨基酸甲酯或二肽甲酯缩合,制备了5-氟尿嘧啶N1-甲酰基二肽甲酯和三肽甲酯。5-氟尿嘧啶-N1-甲酰基二肽甲酯也可采用5-氟尿嘧啶-N1-甲酰氯与二肽甲酯直接反应制备。  相似文献   

11.
This paper reports an investigation on optimal separation and detection for animal prostaglandins, i.e. prostaglandin E(2)(PGE(2)) and thromboxane B(2)(TXB(2)) by capillary zone electrophoresis with indirect detection at 254 nm. It was found that the addition of 3 mmol l(-1) phosphomolybdic acid in 5 mmol l(-1) chromate buffer can lower the detection limits of PGE(2) and TXB(2), i.e. the calculated detection limits (k=3) with phosphomolybdic acid is 0.05 mug ml(-1), only a 1/5-fold improvement compared to that without phosphomolybdic acid. Also adding phosphomolybdic acid, the mechanism for improving detection limits was explained properly. The separation and detection of PGE(2) and TXB(2) can be completed in 4 min. The linear ranges for PGE(2) and TXB(2) were the same, i.e. 5 approximately 80 mug ml(-1), analytical precision (n=8) was 1.2 approximately 1.4% and 3.5% for the measurement of migration times and determination of peak height, respectively. The approach was demonstrated in the lung tissue of SD rats, the measurement results were in good agreement with previous investigations.  相似文献   

12.
A versatile synthetic method for preparing 4‐hydroxyquinolone and 2‐substituted quinolone compounds from simple benzoic acid derivatives was demonstrated. The synthetic strategies involve the use of well known ethyl acetoacetate synthesis, malonic ester synthesis and reductive cyclization. The key intermediates were keto esters 4a‐e , which could be transformed to 4‐hydroxyquinolones 5a,b or 2‐substituted quinolone ethyl esters 6a‐c depending on the reaction conditions. 4‐Hydroxyquinolone analogues were prepared and investigated for N‐methyl‐D‐aspartate (NMDA) activity in vitro. Among these derivatives, 6,7‐difluoro‐3‐nitro‐4‐hydroxyquinolin‐2(1H)‐one ( 9 ) exhibited moderate activity.  相似文献   

13.
A synthetic protocol for the tert-butyl-substituted dichalcogenoimidodiphosphinates [Na(tmeda){(EPtBu(2))(2)N}] (3 a, E=S; 3 b, E=Se; 3 c, E=Te) has been developed. The one-electron oxidation of the sodium complexes [Na(tmeda){(EPR(2))(2)N}] with iodine produces a series of neutral dimers (EPR(2)NPR(2)E--)(2) (4 b, E=Se, R=iPr; 4 c, E=Te, R=iPr; 5 a, E=S, R=tBu; 5 b, E=Se, R=tBu; 5 c, E=Te, R=tBu). Attempts to prepare 4 a (E=S, R=iPr) in a similar manner produced a mixture including HN(SPiPr(2)). Compounds 4 b, 4 c and 5 a-c were characterised by multinuclear NMR spectra and by X-ray crystallography, which revealed two alternative structures for these dimeric molecules. The derivatives 4 b, 4 c, 5 a and 5 b exhibit acyclic structures with a central chalcogen-chalcogen linkage that is elongated by approximately 2 % (E=S), 6 % (E=Se) and 8 % (E=Te) compared to typical single-bond values. By contrast, 5 c adopts an unique spirocyclic contact ion-pair structure in which a [(TePtBu(2))(2)N](-) ion is Te,Te' chelated to an incipient [(TePtBu(2))(2)N](+) cyclic ion. DFT calculations of the relative energies of the two structural isomers indicate a trend towards increasing stability for the contact ion pair relative to the corresponding dichalcogenide on going from S to Se to Te for both the isopropyl and tert-butyl series. The two-electron oxidation of [Na(tmeda){(EPtBu(2))(2)N}] (E=S, Se, Te) with iodine produced the salts [(EPtBu(2))(2)N](+)X(-) (7 a, E=S, X=I(3); 7 b, E=Se, X=I; 7 c, E=Te, X=I), which were characterised by X-ray crystallography. Compound 7 a exists as a monomeric, ion-separated complex with [d(S--S)=2.084(2) A]; 7 b and 7 c are dimeric [d(Se--Se)=2.502(1) A; d(Te--Te)=2.884(1) A].  相似文献   

14.
The electroreduction of Ar-substituted methyl cinnamates in acetonitrile gave all-trans cyclized hydrodimers stereoselectively (58 approximately 90% de). In all cases, small amounts (<10% yield) of meso hydrodimers were also formed. The electrolysis was performed conveniently using an undivided cell at a constant current. The transition states for the hydrocoupling were calculated with semiempirical methods. The all-trans cyclized hydrodimers were transformed to C(2)-symmetric dl-3,4-diaryladipic acids and trans-3,4-diarylcyclopentanones. The chiral auxiliary [(1R)-exo]-3-exo-(diphenylmethyl)borneol, prepared from (1R)-(+)-camphor, was highly effective for the stereoselective hydrocoupling of its cinnamates by electroreduction. From the resulting hydrodimers, (3R,4R)-3,4-diaryladipic acid esters and (3R,4R)-3,4-diarylhexane-1,6-diols were synthesized in 87-95% ee.  相似文献   

15.
Low picogram levels of the E series prostaglandins, PGE1, PGE2, 19-hydroxy PGE1 and 19-hydroxy PGE2, in human semen were analysed by silica capillary column gas chromatography with electron-capture detection after conversion to the methyl ester O-trimethylsilyl derivatives of the corresponding B series prostaglandins. The method was used to detect traces of semen on post-coital vaginal swabs, and on rectal, oral and skin swabs after simulated sexual acts. Semen was detectable on a vaginal swab taken 58 h after intercourse, and was readily detectable for at least 6 h on rectal and skin swabs. Preliminary results suggest that the ratios of prostaglandins on vaginal swabs may indicate how recently intercourse occurred.  相似文献   

16.
The [4 + 2] cycloadditions of 2-oxobut-3-enenitrile ( 1a ), 2-oxopent-3-enenitrile ( 1b ), and ethyl 4-cyano-4-oxobut-2-enoate ( 1c ) with 1,3-dimethyluracil ( 2 ), 1,3, 6-trimethyluracil ( 9 ), or 1,3,5-trimethyluracil ( 16 ) were investigated. The reactions of 1a with 2 or with 9 lead to bicyclic adducts 3 and 10 , respectively. These hexahydro-cis-pyranopyrimidines undergo ring opening under acidic conditions, restoring in 4 and 11 , respectively, an uracil system comprising 2-hydroxybut-2-enenitrile as a side chain at C(5). The surprisingly stable enols tautomerize slowly to the corresponding acyl cyanides 6a and 13a , respectively. Reacting 1b or 1c with 2 and with 9 does not afford cycloadducts; instead the uracil derivatives 6b, c and 13b, c , respectively, show up, carrying at C(5) α-oxobutanenitrile side chains. Cleavage of the acyl cyanide functions in 6a–c and 13a–c with nucleophilic agents produces various acids, esters, or amides, i.e. derivatives 8a–c and 15–c , respectively. The methyl esters 8a (X ? MeO, R ? H) and 15a (X ? MeO, R ? H) are also formed directly from the adducts 3 and 10 , respectively, with acid or base catalysis in presence of MeOH. The cycloadducts 17a and 17c , resulting from the reaction of 1a and 1c with 16 , respectively, have a Me group at the ring junction C(4a) and are stable. The structure of 17c proves that this hetero-Diels-Alder addition of inverse electron demand follows the endo-mode.  相似文献   

17.
Methyl- and phenyl-substituted N-(ethoxycarbonyl)-2-azabicyclo[2.2.0]hex-5-enes 6 have been prepared by photoirradiation of appropriately substituted 1,2-dihydropyridines. Torquoselectivity is observed in the synthesis of the 3-endo-methyl- and 3-endo-phenyl-2-azabicyclo[2.2.0]hexenes 6c-e from 2-methyl- and 2-phenyl-1,2-dihydropyridines 5c-e. Products formed upon addition of bromine to 3-endo-, 4-, and 5-methyl- and 3-endo-phenyl-substituted N-(ethoxycarbonyl)-2-azabicyclo[2.2.0]hex-5-enes 6a-f were substituent dependent. For 6a,b, which lack substituents at C(3) or C(5), mixtures of unrearranged dibromides 8a,b and rearranged dibromides 9a,b were obtained. With the 3-endo-substituents in 6c-e, only rearranged dibromides 9c-e were formed; 5-methyl substitution afforded mainly unrearranged dibromide 8f and some allylic bromide 10. Both unrearranged 5-endo,6-exo-dibromo-2-azabicyclo[2.2.0]hexanes 8 and rearranged 5-anti-6-anti-dibromo-2-azabicyclo[2.1.1]hexanes 9 are formed stereoselectively. The dibromoazabicyclo[2.1.1]hexanes 9 have been reductively debrominated to afford the first reported 2-azabicyclo[2.1.1]hexanes 11 with alkyl or aryl substituents at C-3.  相似文献   

18.
The first derivatives of catenated cyclotetraphosphinophosphonium cations, [(PhP)4PPhMe]+ (8a), [(MeP)4PMe2]+ (8b), [(CyP)4PPh2]+ (8d), [(CyP)4PMe2]+ (8e), [(PhP)4PPh2]+ (8f), [(PhP)4PMe2]+ (8g), are synthesized as trifluoromethanesulfonate (triflate, OSO2CF3-) salts through the reaction of cyclopentaphosphines (PhP)5 (4a) or (MeP)5 (4b) with methyl triflate (MeOTf) or by a net phosphenium ion [PR2+, R = Ph, Me; from R2PCl and trimethylsilyltriflate (Me3SiOTf)] insertion into the P-P bond of either cyclotetraphosphine (CyP)4 (3c) or cyclopentaphosphines (PhP)5 (4a) or (MeP)5 (4b). Although more conveniently prepared from 4a, compound 8a[OTf] can also be formed from (PhP)4 (3a) and MeOTf, and derivatives 8f[OTf] and 8g[OTf] are also accessible through reactions of 3a and R2PCl/Me3SiOTf with R = Ph or Me, respectively. A tetrachlorogallate salt of [(PhP)4PPhtBu]+ (8c) has been synthesized by alkylation of 4a with tBuCl/GaCl3. 31P[1H] NMR parameters for all derivatives of 8 have been determined by iterative simulation of experimental data. Derivatives 8a[OTf], 8b[OTf], 8c[GaCl4], 8e[OTf], 8f[OTf], and 8g[OTf] and have been characterized by X-ray crystallography, showing the most favorable all-trans configuration of substituents for the phosphine centers, thus minimizing steric interactions. Each derivative adopts a unique envelope or twist conformation of C1 symmetry. The effective C2 symmetry observed for 8b, d, e, f, and g in solution, signified by their 31P[1H] NMR AA'BB'X spin systems, implies a rapid conformational exchange for derivatives of 8. The core frameworks of the cations in the solid state are viewed as snapshots of different conformational isomers within the solution-phase pseudorotation process.  相似文献   

19.
A reaction of methyl (4R,5R)-4,5-epoxy-2(E)-hexenoate 1 with N-benzylmethylamine gave a diastereomerically pure methyl (4R,5R)-4,5-epoxy-(3S)-N-benzylmethylamino hexanoate 6 and methyl (4S,5R)-4-N-benzyl-methylamino-5-hydroxy-2(E)-hexenoate 7. The former was chemoenzymatically converted to (-)-osmundalactone 11, which is an aglycone of osmundalin. On the other hand, the directly conjugated addition of dimethylamine to methyl (4S,5S)-4,5-epoxy-2(E)-hexenoate 1 followed by treatment with MeOH at 40 degrees C exclusively provided methyl (4R,5S)-4-dimethylamino-5-hydroxy-2(E)-hexenoate 16, which was converted into L-(-)-forosamine 18.  相似文献   

20.
Thermospray high-performance liquid chromatography-mass spectrometry (HPLC-MS) can be a powerful tool for characterizing eicosanoids in complex biological samples. The positive ion spectra obtained from primary prostaglandins such as PGE1 PGE2, 19-OHPGE1, 19-OHPGE2, PGF2 alpha, PGD2, 6-keto-PGF1 alpha and from leukotriene B4 are very simple, with base peaks corresponding to ions arising from the loss of H2O from the (M + H)+ and (M + NH4)+ ions, except for PGB2 and PGF2, where the latter two ions predominate. The application of this technique to the concurrent determination of the E1 and E2 prostaglandins and their 19-hydroxylated derivatives in human semen is described. The technique affords a moderate level of sensitivity (5-20 ng on-column) and excellent specificity so that virtually no sample manipulation is required other than dilution in acetone and centrifugation. The clear supernatant is injected directly into the HPLC-MS system. A similar analysis by either gas chromatography (GC) or GC-MS would need multi-step derivatization, thus increasing the sample manipulation required and the total analysis time.  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