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 703 毫秒
1.
Syntheses of 11-acety1-2-phenylpyrimido[5,4-c][1,5]benzoxazepin-5(11H)one ( 16a ) and analogs ( 16b,c, 22 ) were described. The reaction of 4-chloro-2-phenyl-5-pyrimidinecarboxylic acid ethyl ester ( 7 ) with 2-aminophenol afforded 4-(2-hydroxyanilino)-2-phenyl-5-pyrimidine-carboxylic acid ethyl ester ( 8a ). The latter was also prepared by catalytic reduction of 4-(2-nitrophenoxy)-2-phenyl-5-pyrimidinecarboxylic acid ethyl ester ( 9 ), which was obtained from 7 and 2-nitrophenol. Involvement of 4-(2-aminophenoxy)-2-phenyl-5-pyrimidinecarboxylic acid ethyl ester ( 12a ) in this reduction as an intermediate was demonstrated by an independent synthesis of 12a and its subsequent rearrangement to 8a. Hydrolysis of 8a or 12a gave 4-(2-hydroxyanilino)-2-phenyl-5-pyrimidinecarboxylic acid ( 15a ). Reaction of 15a with acetic anhydride afforded 16a , the first member of a novel ring system, the pyrimido[5,4- c ][1,5]-benzoxazepin. Additional examples ( 16b,c ) were prepared similarly. The corresponding 11-ethyl derivative ( 22 ) was prepared in similar fashion, starting with 7 and 2-ethylaminophenol. A possible reaction mechanism for the formation of 16a-c from 15a-c and acetic anhydride was discussed.  相似文献   

2.
[structure: see text] An ester dienolate [2,3]-Wittig rearrangement was utilized to access the alkylated citric acid skeleton 6 that is characteristic for the viridiofungins and other members of the alkyl citrate family of secondary natural products. The [2,3]-sigmatropic rearrangement of (Z,Z)-15 provided the rearrangement product (+/-)-syn-16 in moderate yield and with very good diastereoselectivity. A Julia-Kocienski olefination efficiently served to connect the polar head (+/-)-syn-26 with the lipophilic tail (32a-c) of the viridiofungins. Amide formation between the racemic viridiofungin precursors 35a-c and the enantiomerically pure amino acid L-tyrosine methyl ester followed by preparative reversed-phase HPLC provided the isopropyl dimethyl ester of viridiofungin A ((+)-39a), A2 ((+)-39b), and A4 ((+)-39c) as well as the nonnatural diastereomers (-)-38a-c.  相似文献   

3.
3-Phenylpyrazole-5-diazonium chloride (5) couples with benzenesulfonylacetone (9a) , benzenesulfonylacetophenone (9b) , ethyl benzenesulfonylacetate 9c , and ethyl benzoylacetate (12b) in ethanol in the presence of sodium acetate at room temperature to afford the pyrazolo[3,2-c]-1,2,4-triazine derivatives 11a and 11b and the acyclic hydrazones 10c and 13 respectively. The products 11a,b and 10c can also be obtained from the reaction of the corresponding hydrazidoyl halides 8a-c with sodium benzenesulfinate in high yield. The hydrazones 10c and 13 can be cyclised thermally or under the influence of acid into pyrazolo[3,2-c]-1,2,4-triazine derivatives 11c and 14 respectively.  相似文献   

4.
The action of ammonium acetate on 5-arylidene-3-phenyl-2-methylmercaptohydantoins 1g,h in acetic acid led to the formation of the 5-arylidene-3-phenylhydantoin derivatives 4a,b . In absence of a solvent, ring opening and rearrangement took place with the formation of the 5-arylidene-N2-phenylglycocyamidine derivatives 7a-c . Compounds 7a-c reacted with methyl iodide to afford the corresponding 3-methyl derivatives 9a-c . The structures of the synthesised products were established and the mechanism proposed for the rearrangement reaction was discussed.  相似文献   

5.
The trimethylsilyl-protected enynes 9a-c and 14a,b with alkynyl substituents on the three-membered ring or on the double bond of a methylenecyclopropane or a bicyclopropylidene moiety were prepared in two steps from the alcohols 6a-c and 12a,b, respectively, by conversion to the iodides and their coupling with lithium (trimethylsilyl)acetylide (8) in 38-73% overall yields. The bicyclopropylidene derivative 9d was synthesized in 49% yield directly from bicyclopropylidene (3) by lithiation followed by coupling with (5-iodopent-1-ynyl)trimethylsilane (11). Enynes 9b-d were protiodesilylated by treatment with K2CO3 in methanol to give the corresponding unprotected enynes 10b-d in 53, 74 and 94% yield, respectively. Enynes 17a-c with a carbonyl group adjacent to the acetylenic moiety were synthesized from oxo derivatives 15a-c by Wittig olefination followed by coupling with 8 in 47, 18 and 12% overall yield, respectively. Pauson-Khand reactions of the methylenecyclopropane derivatives with a substituent on the ring (9a,b and 10a) as well as on the double bond (14a,b and their in situ prepared protiodesilylated analogues) proceeded smoothly by stirring of the corresponding enyne with [Co2(CO)8] in dichloromethane at ambient temperature followed by treatment of the formed complexes with trimethylamine N-oxide under an oxygen atmosphere at -78 degrees C to give tricyclic or spirocyclopropanated bicyclic enones 18a,b, 19a, 20a,b, 21a,b in good yields. Alkynylbicyclopropylidene derivatives 9c,d and 10c,d formed the corresponding cobalt complexes at -78 to -20 degrees C. Treatment of the latter with N-methylmorpholine N-oxide under an argon atmosphere at -20 degrees C gave the spirocyclopropanated tricyclic enones 18c, 19c and 18d in 31-45% yields. The structure of 19c was proved by X-ray crystal structure analysis. The cyclization of enynones 17a-c in MeCN at 80 degrees C gave the spirocyclopropanated bicyclic diketones 22a-c in 38-65% yields. Intramolecular PKRs of the enynes 25a,d with a chiral auxiliary adjacent to the triple bond gave the corresponding products 26a,d in 70 and 79% yield, respectively, as 5:1 and 8:1 mixtures of diastereomers, respectively. Addition of lithium dimethylcuprate or higher order cuprates to the double bond of the former furnished bridgehead-substituted bicyclo[3.3.0]octanones 27a-c in 57-86% yields. Protiodesilylation of 27a followed by acetal cleavage gave the enantiomerically pure spirocyclopropanated bicyclo[3.3.0]octanedione (1R,5R)- 29a with [alpha]D(20)=-148 (c=1.0 in CHCl3) in 55% overall yield.  相似文献   

6.
Reaction of 2-(3-,4-)pyridinecarboxaldehydes 5 with carbomethoxymethylene triphenylphosphorane afforded predominantly E-methyl-3-(pyridinyl)-2-propenoates 7. Oxidation of 7a-c with m-chloroperbenzoic acid gave methyl E-3-(1-oxidopyridinyl)-2-propenoates 8a-c in high yield. The Darzen's reaction of 5a-c with methyl bromoacetate yielded a mixture of stereoisomers cis- 9a-c and methyl trans-3-(pyridinyl)-2,3-epoxy-propanoates 10a-c in a ratio of 2:1. Oxidation of cis- 9a-c and trans- 10a-c afforded the corresponding cis- 11a-c and methyl trans-3-(1-oxidopyridinyl)-2,3-epoxypropanoates 12a-c in good yield. The reaction of 11a and 12a with cyclic amines as piperidine gave the respective threo- 13 and methyl erythro-2-(1-piperidino)-3-hydroxy-3-(1-oxido-2-pyridino)propanoate 14. The sodium borohydride reduction of the N-alkoxylcarbonyl pyridinium salts of 9c and 10c afforded the corresponding N-alkoxycarbonyl-1,2-dihydropyridyl derivatives 15 and 16. A number of selected compounds ( 7a-c , 9a-c , 10a , 10c , 11a-c and 12a , 12c ) were found to be inactive in the P388 Lymphocytic screen. Compounds 9-12 did not react with the model nucleophile ethanethiol in phosphate buffer at 37°.  相似文献   

7.
A highly chemoselective sequence of Stille and Heck couplings on the heterocyclic bromoenol triflates 2 a-c with the bicycloalkenylstannanes cis-3 and trans-3 furnished the intermediate bromobutadienes 4 a-c in good yields ranging from 73-94 %. A modified Heck coupling protocol employing the palladacycle 8 and an additional bidentate ligand such as 1,4-bis(diphenylphosphinyl)butane allowed a significant reduction in catalyst loading while still obtaining the heterocyclic 1,3,5-hexatrienes 5 a-c in good yields (71-94 %). The unsymmetrically substituted 1,3,5-hexatrienes 5 a-c in solution underwent 6pi-electrocyclizations following an optimized microwave-heating protocol to yield the steroidal tetracycles cis-7 a-c and trans-7 b (59-69 %). Tetracycles cis-7 a-c are the products of a subsequent 1,5-hydrogen shift to the thermodynamically more stable, more highly substituted diene units. Removal of the tert-butyl groups provided the novel steroidal delta-amino acid 9 a and the delta-amino acid derivatives 9 b, c in good yields (76-86 %).  相似文献   

8.
The reaction of 6-chloro-2-(1-methylhydrazino)quinoxaline 4-oxide 5 with a 2-fold molar amount of ethyl chloroglyoxalate gave ethyl 8-chloro-4-methyl-4H-1,3,4-oxadiazino[5,6-b]quinoxaline-2-carboxylate 6 , whose reaction with hydrazine hydrate afforded the C2-hydrazinocarbonyl derivative 7 . The reaction of compound 7 with nitrous acid provided the C2-acylazide derivative 8 , which was converted into the C2-amino 9 , C2-carbamate 11a-c, 12a,b , and C2-ureido 13a-c, 14 derivatives. The mass spectral fragmentation patterns were examined for compounds 10–14 , wherein the molecular ion peak did not appear in the mass spectra of compounds 10c, 11a-c, 12a,b, 13c , and 14.  相似文献   

9.
New alternating poly(amide-ester)s derived from β-hydroxy acids and α-amino acids 3a,b or ϵ-aminocaproic acid 4a-c were prepared. Two approaches were considered: (i) polycondensation of N-(β-hydroxyacyl)-amino acids 1a,b and 2b,c and (ii) ring-opening polymerization of cyclic amide-esters 5a-c and 6a-c . For all the linear precursors polycondensation reactions result in oligomers with number average molecular weights lower than 5000. The ring-opening polymerization of the cyclic precursors is substrate specific and is sensitive to changes in the polymerization conditions. For N-(3-hydroxybutyroyl)-ϵ-aminocaproic acid lactone [c(3HB-ϵAC); 5b ] (IUPAC nomenclature: 2-methyl-5-aza-1-oxa-cycloundecan-4,11-dione) bulk and solution polymerizations result in oligomers with an alternating ester amide microstructure. Polymerization of N-(3-hydroxypropionyl)-ϵ-aminocaproic acid lactone [c(3HP-ϵAC); 5a ] (IUPAC nomenclature: 5-aza-1-oxa-cycloundecan-4-11-dione) in dimethylformamide solution and with Bu2Sn(OMe)2 as initiator high molecular weight linear, semi-crystalline polymers were obtained (Tm = 145.9°C). Polymerization of N-(hydroxypivaloyl)-ϵ-aminocaproic acid lactone [c(HPv-ϵAC); 5c ] (IUPAC nomenclature: 3,3-dimethyl-5-aza-1-oxa-cycloundecan-4-11-dione) in bulk results in amorphous alternating poly(amide-ester)s with cyclic structure (Tg = 6.8°C). The fourteen membered cyclo(diamide-diester)s 6a-c (IUPAC nomenclatures:: 4,11-diaza-1,8-dioxa-cyclotetradecan-2,5,9,12-tetraone ( 6a ), 7,14 dimethyl-4,11-diaza-1,8-dioxa-cyclotetradecan-2,5,9,12-tetraone ( 6b ), 3,10-dimethyl-4,11-diaza-1,8-dioxa-cyclotetradecan-2,5,9,12-tetraone ( 6c ) based on β-hydroxy acids and α-aminoacids could not be polymerized.  相似文献   

10.
3-Oxo-3-phenyl-2-(p-tolylhydrazono)propanal (1a) undergoes condensation with ethyl cyanoacetate in acetic acid in the presence of ammonium acetate to yield either 2-hydroxy-6-phenyl-5-p-tolylazonicotinic acid ethyl ester (6a) or 2-amino-6-phenyl-5-ptolyl-azonicotinic acid ethyl ester (8), depending on the reaction conditions. Similarly, other 3-oxo-3-aryl-2-arylhydrazonopropanals 1a,b condense with active methylene nitriles 2c,d to yield arylazonicotinates 6b,c. In contrast, 2-[(4-nitrophenyl)-hydrazono]-3-oxo-3-phenyl-propanal (1c) reacts with ethyl cyanoacetate to yield ethyl 6-(4-nitrophenyl)-2-oxo-2,6-dihydropyrido[3,2–c]cinnoline-3-carboxylate (11), via a novel 6π-electrocyclization pathway. Finally, 3-oxo-2-(phenylhydrazono)-3-p-tolylpropanal (1d) condenses with 2a-c to yield pyridazinones 13a-c.  相似文献   

11.
Irradiation of 1-substituted benzotriazole arylhydrazones 3a-c, 4a,b and 5a,b with a 16 W low pressure mercury arc-lamp (254 nm) for 24 h gave phenanthridin-6-yl-2-phenyldiazines 9a-c, phenanthridin-6(5H)-ones 10a-c, 1-anilinobenzimidazoles 11a-c, 2-aryl-1H-benzimidazoles 12a-c, 1-arylamino-1H-benzimidazol-2-carboxylic acid ethyl esters 14a,b, 1-aryl-1H, 9H-benzo [4,5][1,2,3] triazolo[1,2-a]tetrazole-3-carboxylic acid ethyl esters 16a,b, 1-arylamino-2-benzoylbenzimidazoles 18a,b and 2-benzoylbenzoxazole 21.  相似文献   

12.
The aminodihydropentalene derivative 1a reacts with the Lewis acidic RB(C(6)F(5))(2) boranes (2a-c) by C-C bond cleavage to yield the formal borylene insertion products 3. In contrast, 1a,b react with HB(C(6)F(5))(2) at 55 °C by elimination of dihydrogen to yield the iminium-stabilized zwitterionic heterofulvenes 10a,b. The reaction pathways were studied by preparation of the kinetically controlled intermediates 7a,b and the thermodynamically controlled products 9a,b, monitored by variable-temperature NMR experiments, and supported by DFT calculations. The trapping reactions of 9a with HCl and PhCHO, respectively, led to the addition products 13 and 14. Compounds 3c, 7a,b, 10a,b, 11, 13, and 14 were characterized by X-ray diffraction.  相似文献   

13.
Regioselective 1,3-dipolar cycloaddition of nitrile oxides 5a-c to ethyl o-hydroxycinnamate (3) gave the corresponding ethyl trans-3-aryl-4,5-dihydro-5-(2-hydroxyphenyl)-4-isoxazolecarboxylates 6a-c . Their structure was confirmed by reductive cleavage to 1 and compounds 9a-c . Compounds 6a-c afforded upon heating in the presence of pyridine the 3-aryl-4H-[1]benzopyrano[3,4-d]isoxazol-4-ones 11a-c . Compound 10c was also isolated from 6c and transformed thermally into 11c .  相似文献   

14.
Addition of the new phosphonium carborane salts [HPR(3)][closo-CB(11)H(6)X(6)] (R = (i)Pr, Cy, Cyp; X = H 1a-c, X = Br 2a-c; Cy = C(6)H(11), Cyp = C(5)H(9)) to [Rh(nbd)(mu-OMe)](2) under a H(2) atmosphere gives the complexes Rh(PR(3))H(2)(closo-CB(11)H(12)) 3 (R = (i)Pr 3a, Cy 3b, Cyp 3c) and Rh(PR(3))H(2)(closo-CB(11)H(6)Br(6)) 4 (R = (i)Pr 4a, Cy 4b, Cyp 4c). These complexes have been characterised spectroscopically, and for 4b by single crystal X-ray crystallography. These data show that the {Rh(PR(3))H(2)}(+) fragment is interacting with the lower hemisphere of the [closo-CB(11)H(6)X(6)](-) anion on the NMR timescale, through three Rh-H-B or Rh-Br interactions for complexes 3 and 4 respectively. The metal fragment is fluxional over the lower surface of the cage anion, and mechanisms for this process are discussed. Complexes 3a-c are only stable under an atmosphere of H(2). Removing this, or placing under a vacuum, results in H(2) loss and the formation of the dimer species Rh(2)(PR(3))(2)(closo-CB(11)H(12))(2) 5a (R = (i)Pr), 5b (R = Cy), 5c (R = Cyp). These dimers have been characterised spectroscopically and for 5b by X-ray diffraction. The solid state structure shows a dimer with two closely associated carborane monoanions surrounding a [Rh(2)(PCy(3))(2)](2+) core. One carborane interacts with the metal core through three Rh-H-B bonds, while the other interacts through two Rh-H-B bonds and a direct Rh-B link. The electronic structure of this molecule is best described as having a dative Rh(I) --> Rh(III), d(8)--> d(6), interaction and a formal electron count of 16 and 18 electrons for the two rhodium centres respectively. Addition of H(2) to complexes 5a-c regenerate 3a-c. Addition of alkene (ethene or 1-hexene) to 5a-c or 3a-c results in dehydrogenative borylation, with 1, 2, and 3-B-vinyl substituted cages observed by ESI-MS: [closo-(RHC[double bond, length as m-dash]CH)(x)CB(11)H(12-x)](-)x = 1-3, R = H, C(4)H(9). Addition of H(2) to this mixture converts the B-vinyl groups to B-ethyl; while sequential addition of 4 cycles of ethene (excess) and H(2) to CH(2)Cl(2) solutions of 5a-c results in multiple substitution of the cage (as measured by ESI-MS), with an approximately Gaussian distribution between 3 and 9 substitutions. Compositionally pure material was not obtained. Complexes 4a-c do not lose H(2). Addition of tert-butylethene (tbe) to 4a gives the new complex Rh(P(i)Pr(3))(eta(2)-H(2)C=CH(t)Bu)(closo-CB(11)H(6)Br(6)) 6, characterised spectroscopically and by X-ray diffraction, which show coordination of the alkene ligand and bidentate coordination of the [closo-CB(11)H(6)Br(6)](-) anion. By contrast, addition of tbe to 4b or 4c results in transfer dehydrogenation to give the rhodium complexes Rh{PCy(2)(eta(2)-C(6)H(9))}(closo-CB(11)H(6)Br(6)) 7 and Rh{PCyp(2)(eta(2)-C(5)H(7))}(closo-CB(11)H(6)Br(6)) 9, which contain phosphine-alkene ligands. Complex has been characterised crystallographically.  相似文献   

15.
A convenient approach to phenyl (3-pyridazinyl) ketone ( 6 ) and phenyl(3-pyridazinyl)methanol ( 7 ) is proposed. Reactions of the related diarylmethyl chloride 8 with various N- and S-nucleophiles were found to afford the expected amines 9a-c, 10a-c and thioethers 11a,b in satisfactory yields.  相似文献   

16.
吡喃酮是许多天然产物的结构单元,我们曾由4-异丁酰基庚二酸在过量醋酸酐及乙酰氯存在下回流得到7-氧代-8,8-二甲基-△~9-六氢香豆素.本文由二氰乙基-β-二酮进行酮解水解反应得到4-酰基庚二酸1_(a-c)。 在过量醋酸酐、乙酰氯存在下由1_a、1_c为底物进行反应没有得到双环的香豆素衍生物.其产物和单纯以乙酐为缩合剂时的产物2_a、2_c相同,产率分别为68%、63%。2_c可在硫酸铁催化  相似文献   

17.
Reactions of dibenzo[b,e]thiepin-11(6H)-one (4) with 2-, 3- and 4-(dimethylaminomethyl)phenylmagnesium bromide afforded the tertiary alcohols 5a,b,c. The aldehydes 7 and 8 gave similarly the secondary alcohols 9a,b,c and 10c . Numerous attempts to prepare the corresponding ketones, especially by oxidation of 9a,b,c and 10c were unsuccessful. Only the oxidation of 9c with tetrabutylammonium chromate in chloroform afforded the desired ketone 16 . Its formation was accompanied by an important side reaction consisting in a cleavage of the “retro-ene-reaction” type leading to compound 11 and the aldehyde 13c which reacted with the chloroform present to give the alcohol 17 . Compounds 5a,b,c, 9a,b,c and 16 were tested as potential antidepressants but with the exception of some effects in the test of potentiation of yohimbine toxicity in mice, they proved inactive in this line.  相似文献   

18.
4,5-Diaryl-1H-pyrazole-3-ol was utilized as a versatile template to synthesize several classes of compounds such as pyrazolo-oxazines 7, pyrazolo-benzooxazines 9, pyrazolo-oxazoles 10, and its analogues 11a-c as potential COX-2 inhibitors. Compounds 11b,c were successfully synthesized with use of pyridinium p-toluenesulfonate mediated cyclization of the ketal intermediate. Diaryl-pyrazolo-benzooxazepine analogues were synthesized by using Cu-mediated cyclization of the O-alkylated arylbromide intermediate. Arylsulfonamides were synthesized efficiently on a large scale with 4-[4-(4-fluorophenyl)-5-hydroxy-2H-pyrazol-3-yl]benzenesulfonamide 31 template readily synthesized from commercially available 4-sulfamoyl benzoic acid 29. The structure of a representative compound from each class was confirmed by X-ray crystallography. Selected compounds tested for inhibitory activity against COX-1 and COX-2 enzymes showed good selectivity for COX-2 versus COX-1 enzyme.  相似文献   

19.
1-Benzyl-3-cyanopyrrole-2-carbonyl azide (5) underwent a Curtius rearrangement followed by quenching with alcohols to form the corresponding carbamates (6a-c). The carbamates 6a,b were unblocked to give the desired 2-amino-1-benzyl-3-cyanopyrrole (1a). A more facile procedure was subsequently developed for the synthesis of 1-substituted 2-amino-3-cyanopyrroles. N-Substituted aminoacetaldehyde dimethyl acetals (7a-c) were condensed with malononitrile in the presence of p-toluenesulfonic acid monohydrate to afford the corresponding 1-substituted 2-amino-3-cyanopyrroles (1a-c).  相似文献   

20.
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   

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