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 46 毫秒
1.
1,2,3,2',3',4',6'-Hepta-O-acetyl-beta-lactose (4) was coupled with 2,3,6,2',3',4',6'-hepta-O-acetyl-alpha-lactosyl bromide (7) in the presence of Hg(CN)2 to afford 1,2,3,2',3',4',6'-hepta-O-acetyl-6-O-(2,3,6,2',3',4',6'-hepta-O-acetyl-b eta- lactosyl)-beta-lactose (11) which, upon O-deacetylation, gave 6-O-beta-lactosyl-alpha,beta-lactoses (64% from 4). In contrast, the reaction of 7 with benzyl 2,3,2',3',4',6'-hexa-O-acetyl-beta-lactoside in the presence of Hg(CN)2 produced 3,6,2',3',4',6'-hexa-O-acetyl-1,2-O- (2,3,2',3',4',6'-hexa-O-acetyl-1-O-benzyl-beta-lactos-6-yl orthoacetyl)-alpha-lactose (63%) and 3,6,2',3',4',6'-hexa-O-acetyl-1,2-O-(1- cyanoethylidene)-alpha-lactose (27%). The glycosidation of 4 using 2,3,4,6-tetra-O-acetyl-alpha-D-galactopyranosyl bromide in the presence of Hg(CN)2 afforded, after deprotection, 4,6-di-O-beta-D-galactopyranosyl-alpha,beta-D-glucoses (66%). The reaction of 11 with 1,2-di-O-benzyl-(R,S)-glycerols and trimethylsilyl trifluoromethanesulfonate yielded, after deprotection, 1-O-(6-O-beta-lactosyl-beta-lactosyl)-(R,S)-glycerols (18%). Under the same coupling conditions 11 reacted with 2-O-benzylglycerol to form 3-O-acetyl-2-O-benzyl-1-O-[2',3',4',6'-hexa-O-acetyl-6-O-(2,3,6,2',3',4' ,6'- hepta-O-acetyl-beta-lactosyl)-beta-lactosyl]-(R,S)-glycerols (16%).  相似文献   

2.
The regio- and stereospecific oxidation of fluorene, dibenzofuran, and dibenzothiophene was examined with mutant and recombinant strains expressing naphthalene dioxygenase from Pseudomonas sp. strain NCIB 9816-4. The initial oxidation products were isolated and identified by gas chromatography-mass spectrometry and nuclear magnetic resonance spectrometry. Salicylate-induced cells of Pseudomonas sp. strain 9816/11 and isopropyl-beta-D-thiogalactopyranoside-induced cells of Escherichia coli JM109(DE3)(pDTG141) oxidized fluorene to (+)-(3S,4R)-cis-3,4-dihydroxy-3,4-dihydrofluorene (80 to 90% relative yield; > 95% enantiomeric excess [ee]) and 9-fluorenol (< 10% yield). The same cells oxidized dibenzofuran to (1R,2S)-cis-1,2-dihydroxy-1, 2-dihydrodibenzofuran (60 to 70% yield; > 95% ee) and (3S,4R)-cis-3, 4-dihydroxy-3,4-dihydrodibenzofuran (30 to 40% yield; > 95% ee). Induced cells of both strains, as well as the purified dioxygenase, also oxidized dibenzothiophene to (+)-(1R,2S)-cis-1,2-dihydroxy-1, 2-dihydrodibenzothiophene (84 to 87% yield; > 95% ee) and dibenzothiophene sulfoxide (< 15% yield). The major reaction catalyzed by naphthalene dioxygenase with each substrate was stereospecific dihydroxylation in which the cis-dihydrodiols were of identical regiochemistry and of R configuration at the benzylic center adjacent to the bridgehead carbon atom. The regiospecific oxidation of dibenzofuran differed from that of the other substrates in that a significant amount of the minor cis-3,4-dihydrodiol regioisomer was formed. The results indicate that although the absolute stereochemistry of the cis-diene diols was the same, the nature of the bridging atom or heteroatom influenced the regiospecificity of the reactions catalyzed by naphthalene dioxygenase.  相似文献   

3.
Further lead optimization efforts on previously described 1,2,3,4,10,10a-hexahydro-1H-pyrazino[1,2-a]indoles led to the new class of 5,5a,6,7,8,9-hexahydro-pyrido[3',2':4,5]pyrrolo[1,2-a]pyrazines culminating in the discovery of (5aR,9R)-2-[(cyclopropylmethoxy)methyl]-5,5a,6,7,8,9-hexahydro-9-methyl-pyrido[3', 2':4,5]pyrrolo[1,2-a]pyrazine 18 as a potent, full 5-HT(2C) receptor agonist with an outstanding selectivity profile and excellent hERG and phospholipidosis properties.  相似文献   

4.
C(2)-Symmetric chiral diethoxyphosphoramide 4, diethoxythiophosphoramide 5, and diisopropoxyphosphoramide 6 of (1R, 2R)-1,2-diaminocyclohexane were prepared by the reactions of diethoxyphosphinic chloride, diethoxythiophosphinic chloride, and diisopropoxyphosphinic chloride with (1R, 2R)-1,2-diaminocyclohexane, respectively. They were used as catalytic chiral ligands in the asymmetric addition reactions of diethylzinc to aldehydes in the presence of titanium(IV) isopropoxide to give the corresponding sec-alcohols with 43-70% ee. Chiral ligands 4 and 5 gave the sec-alcohols with opposite absolute configuration.  相似文献   

5.
The active catalysts for the enantioselective ring opening (ARO) of meso-stilbene oxide, cis-butene oxide, cyclohexene oxide, cyclopentene oxide, and cyclooctene oxide with various substituted anilines were generated in situ by the reaction of Ti(O(i)Pr)(4) with poly-[(R,R)-N,N'-bis-{3-(1,1-dimethylethyl)-5-methylene salicylidene} cyclohexane-1,2-diamine]-1 and (1R,2R)-N,N'-bis[3,5-di(tert-butyl)salicylidene] cyclohexane-1,2-diamine-2. These catalysts in the presence of nonracemic imine as an additive provided β-amino alcohol in excellent yield (99%) and chiral purity (enantiomeric excess (ee) up to 99%) for the ARO of meso-stilbene oxide with aniline. The same protocol was less effective for the ARO of cyclic epoxides; however, when triphenylphosphine was used as an additive, there was a significant improvement in catalyst performance for the ARO of cyclohexene oxide (yield, 85-90%; ee, 63-67%). Both in situ generated polymeric and monomeric catalysts performed in a similar manner except that the polymeric catalyst Ti(IV)-1 was more active and recycled several times with retention of enantioselectivity when compared with the monomeric catalyst Ti(IV)-2, which was nonrecyclable.  相似文献   

6.
Structure-activity relationships of 1'-acetoxychavicol acetate (ACA) for apoptotic activity against human leukemia HL-60 cells were investigated using optically active ACA and various racemic ACA analogues. Natural-type (or with different acyl group) ACA showed a high apoptotic activity, but the ortho or meta isomers, 4-deacetoxy analogue, and the 2'-3' dehydrogenated derivative had no effect, or a weak activity. Optically active (R)- and (S)-ACA were prepared by a lipase-catalyzed esterification. Using a mixture of vinyl acetate-tetrahydrofuran (1:1 v/v) as a solvent at refluxing temperature, optically pure (R)- and (S)-ACA were obtained (99.7% ee and 99.1% ee, respectively). The apoptosis-inducing effects of both enantiomers were compared by means of an MTT assay and the detection of typical apoptotic phenomena (DNA fragmentation, caspase-3 activation, and PARP cleavage) and these two activities were almost equal. These results indicate that the essential moieties of ACA for apoptotic activity against HL-60 cells are both the presence of a 4-acetoxyl group and an unsaturated double bond between C-2' and C-3', and that the configuration at the 1'-position is unrelated to activity.  相似文献   

7.
The hitherto unknown (-)- and (+)-1-benzylcyclohexan-1,2-diamine hydrochlorides 4a. HCl and 4b. HCl were synthesized by means of diastereoselective alpha-iminoamine rearrangement with subsequent imine reduction and hydrogenolysis. The relative trans-configuration as well as the absolute (1S,2R) and (1R,2S) configurations of 4a and 4b, respectively, were elucidated on the basis of an X-ray analysis of 3b. HCl. The enantiomeric excess (ee) values of the title compounds (>99%) were determined by chiral HPLC on a Chirex (D) Penicillamine column.  相似文献   

8.
Trityl-glycidol was synthesized by in situ derivatization of glycidol, which was prepared by the catalytic asymmetric epoxidation of allyl alcohol. Depending on the enantiomer of diisopropyl tartrate used with the titanium catalyst, either (R)- or (S)-trityl-glycidol was obtained in a "one pot" synthesis in about 50% overall yield. The optical purity, determined by NMR spectroscopy of a Mosher ester, was greater than 98% ee. Nucleophilic opening of the chiral epoxide with dodecyl mercaptan gave optically active 1-S-dodecyl-3-O-trityl-1-thio-glycerol, which was used to synthesize 1-S-dodecyl-2-O-decanoyl-thio-sn-glycero-3-phosphocholine. Opening of the epoxide with methyl xanthate gave a 1,2-trithiocarbonate derivative of trityl glycerol which can be used to synthesize 1,2-bis(S-decanoyl)-1,2-dithio-sn-glycero-3-phosphocholine. Opening of the epoxide with thiodecanoic acid gave 1-S-decanoyl-3-O-trityl-1-thio-glycerol which was used to synthesize 1-S-decanoyl-2-O-decanoyl-1-thio-sn-glycero-3-phosphocholine.  相似文献   

9.
(1R,2S)-1-(3'-Chloro-4'-methoxyphenyl)-1,2propanediol (Trametol, 3), a metabolite of the fungus Trametes sp. IVP-F640 and Bjerkandera sp. BOS55, was synthesized by employing Sharpless asymmetric dihydroxylation as the key step. Similarly, the (1R,2S)-isomer of 1-(3',5'-dichloro-4'-methoxyphenyl)-1,2-propanediol (4), another metabolite of Bjerkandera sp. BOS55, was synthesized by asymmetric dihydroxylation.  相似文献   

10.
Epoxide hydrolase activity of Streptomyces strains   总被引:2,自引:0,他引:2  
The discovery of epoxide hydrolases within a Streptomyces sp. strain collection is described. Screening was performed in 96 well microtiter plates using a modified 4-(p-nitrobenzyl)pyridine assay with styrene oxide, 1,2-epoxy-hexane or 3-phenyl ethylglycidate (3-PEG) as substrates. Out of 120 strains investigated, S. antibioticus Tü4, S. arenae Tü495 and S. fradiae Tü27 exhibited epoxide hydrolase activity. These strains were further investigated by performing laboratory-scale biotransformations utilizing styrene oxide, 1,2-epoxy-hexane and 3-PEG followed by subsequent quantitative analysis employing chiral gas chromatography. The highest conversions were achieved with whole cells from S. antibioticus Tü4 in the presence of 10% (v/v) DMSO. However, enantioselectivity was only satisfying (E = 31) in the presence of 5% (v/v) acetone, which allowed isolation of optically pure non-hydrolyzed (R)-styrene oxide (99% enantiomeric excess (ee)) and (S)-phenyl-1,2-ethandiol (72% ee) at 55% conversion after 24 h. The resolution of 3-PEG proceeded with slightly lower enantioselectivity albeit higher reaction rates. With S. fradiae Tü27 and S. arenae Tü495 enantioselectivity towards styrene oxide was only E = 3-4.  相似文献   

11.
The substrate oxidation profiles of Sphingomonas yanoikuyae B1 biphenyl-2,3-dioxygenase and cis-biphenyl dihydrodiol dehydrogenase activities were examined with 1,2-dihydronaphthalene and various cis-diols as substrates. m-Xylene-induced cells of strain B1 oxidized 1,2-dihydronaphthalene to (-)-(1R,2S)-cis-1,2-dihydroxy-1,2-3,4-tetrahydronaphthalene as the major product (73% relative yield). Small amounts of (+)-(R)-2-hydroxy-1,2-dihydronaphthalene (15%), naphthalene (6%), and alpha-tetralone (6%) were also formed. Strain B8/36, which lacks an active cis-biphenyl dihydrodiol dehydrogenase, formed (+)-(1R,2S)-cis-1,2-dihydroxy-1,2-dihydronaphthalene (51%), in addition to (-)-(1R,2S)-cis-1,2-dihydroxy-1,2,3,4-tetrahydronaphthalene (44%) and (+)-(R)-2-hydroxy-1,2-dihydronaphthalene (5%). The cis-biphenyl dihydrodiol dehydrogenase of strain B1 oxidized both enantiomers of cis-1,2-dihydroxy-1,2-dihydronaphthalene, but only the (+)-(1S,2R)-enantiomers of cis-1,2-dihydroxy-1,2,3,4-tetrahydronaphthalene and cis-1,2-dihydroxy-3-phenylcyclohexa-3,5-diene. The results show that biphenyl dioxygenase expressed by S. yanoikuyae catalyzes dioxygenation, monooxygenation, and desaturation reactions with 1,2-dihydronaphthalene as the substrate, and cis-biphenyl dihydrodiol dehydrogenase catalyzes the enantioselective dehydrogenation of (+)-(1S,2R)-cis-1,2-dihydroxy-1,2,3,4-tetrahydronaphthalene and (+)-(1S,2R)-cis-1,2-dihydroxy-3-phenylcyclohexa-3,5-diene.  相似文献   

12.
Starting from 6-(4'-methoxyphenyl)fulvene (1a), 6-(2',4',6'-trimethoxyphenyl)fulvene (1b), or 6-(3',5'-dimethoxyphenyl)fulvene (1c), [1,2-di(cyclopentadienyl)-1,2-di(4'-methoxyphenyl)-ethanediyl] titanium dichloride (2a), [1,2-di(cyclopentadienyl)-1,2-bis(2',4',6'-trimethoxyphenyl)-ethanediyl] titanium dichloride (2b), and [1,2-di(cyclopentadienyl)-1,2-bis(3',5'-dimethoxyphenyl)-ethanediyl] titanium dichloride (2c) were synthesised. When titanocenes 2a-c were tested against pig kidney carcinoma cells (LLC-PK) inhibitory concentrations (IC50) of 2.8 x 10(-4), 3.6 x 10(-4) and 2.1 x 10(-4) M, respectively, were observed.  相似文献   

13.
1. Monoglycosyl monoglyceride, mono-, di-, tri- and tetraglycosyl diglycerides were isolated from rice bran and characterized for their chemical structures. 2. Monoglycosyl monoglycerides were characterized as Gal(beta 1' leads to 3)-1- or 2-monoacyl-sn-glycerol and Glc(beta 1' leads to 3)-1- or 2-monoacyl-sn-glycerol. 3. The structures of monoglycosyl diglyceride were Gal(beta 1' leads to 3)-1,2-diacyl-sn-glycerol and Glc(beta 1' leads to 3)-1,2diacyl-sn-glycerol. Epimeric separation of the galactosyl and glucosyl glycerides was for the first time achieved by thin-layer chromatography. 4. The main diglycosyl diglyceride was shown to be Gal(alpha 1' leads to 6')-Gal(beta 1' leads to 3)-1,2-diacyl-sn-glycerol. 5. The major structure of triglycosyl diglyceride was characterized as Gal(alpha 1' leads to 6')-Gal(alpha 1' leads to 6')-Gal(beta 1' leads to 3)-1,2-diacyl-sn-glycerol. 6. The representative structure of tetraglycosyl diglyceride was for the first time established as Gal(alpha 1' leads to 6')-Gal(alpha 1' leads to 6')-Gal(a-pha 1' leads to 6')-Gal(beta1' leads to 3)-1,2-diacyl-sn-glycerol.  相似文献   

14.
Kinetic resolution of the enantiomers of trans -1-phenylpropene oxide and (2,3-epoxypropyl)benzene was achieved by yeasts from the genus Rhodotorula. The resolution of trans -1-phenylpropene oxide by Rhodotorula glutinis UOFS Y-0123 yielded (1R,2R)-epoxide (ee >98%, yield 30%) and (1R,2S)-diol (ee 95%, yield 40%). The highest enantio- and regioselectivity toward (2,3-epoxypropyl)benzene resided in Rhodotorula sp. UOFS Y-0448 (E = 6.16), yielding (S)-epoxide (ee 64%, yield 33%) and (R)-diol (ee 67%, yield 28%). This confirms the superiority of yeasts from the Basidiomycetes genera in the enantioselective hydrolysis of epoxides from different structural classes.  相似文献   

15.
The configurations of three enantiopure tetrasubstituted alpha-fluoro cyclohexanones (-)-5Ia, (-)-5IIa and (-)-6a were determined by VCD and proved to be (-)-(2S,5R)-5Ia, (-)-(2R,5R)-5IIa, and (-)-(2R,5R)-6a. The VCD study also identified the conformers populated in CDCl3 solution, including higher-energy gas-phase conformers with equatorial fluorine for 5Ia and 5IIa that are stabilized in CDCl3 solution. Used as catalysts for epoxidation of trans olefins (beta-methylstyrene, stilbene, methyl p-methoxy cinnamate) by oxone, it was found that (-)-5Ia is the most efficient for all trans olefins (providing, respectively, 62%, 90% and 66% ee) but that all three ketones provide high ee% with stilbene (78-90% ee). Moreover, the configurations predicted from the stereo outcome of the epoxidation reaction are identical to those determined by VCD.  相似文献   

16.
1. (3RS,6R)-[6-2H1,6-3H1,6-14C], (3RS,6S)-[6-2H1,6-3H1,6-14C] and (3RS)-[6-3H1,6-14C]mevalonolactones were synthesised from R-[2H1,3H1,2-14C], S-[2H1,3H1,2-14C] and [3h1,2-14C]acetic acids respectively. 2. Each mevalonate was converted into cholesterol by a rat liver preparation. 3. Each cholesterol specimen was converted into androsta-1,4-diene-3,17-dione by incubation with Mycobacterium phlei in the presence of 2,2'.dipyridyl. Each specimen of androsta-1,4-diene-3,17-dione was converted into androsta-1,4-dien-3-one-17-ethylene ketail. 4. The samples of androsta-1,4-dien-3-one-17-ethylene ketal were each converted chemically into oestrones in which the methyl group at C-18 is the only carbon atom that originated from C-6 in mevalonolactone. 5. The oestrone from (3RS)-[6-3H1,6-14C]mevalonolactone was oxidised chemically to acetic acid which was converted into p-bromophenacyl acetate and the 3H/14C ratio was measured. 6. There was no overall loss of tritium from the methyl group of acetic acid, as measured by determining the 3H/14C ratios of the p-bromophenacyl esters, when the synthetic and degradative procedures 1 -- 5 were tested with [3H1,2-14C]acetic acid. 7. The oestrones derived from the 6R and 6S-mevalonolactones were oxidised. The chiralities of the resulting acetates were determined by an established procedure whereby the acetates were converted into 2S-malates which were examined for loss of tritium on equilibration with fumarate hydratase. 8. The oestrone from (3RS,6R)-[6-2H1,6-3H1,6-14C]mevalonate gave acetic acid which was converted into 2S-malate that retained 68.6% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was R. 9. The oestrone from (3RS,6S)-E16-2H1,6-3H1,6-14C]mevalonate was oxidised to acetic acid which was converted into 2S-malate that retained 31.9% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was S. 10. There was no overall change in the configuration of a chiral methyl group between C-6 of mevalonate and C-18 of oestrone. It is cncluded that the intramolecular migration of a chiral methyl group from C-15 in 2,3-oxidosqualene to C-13 in lanosterol is stereospecific and occurs with overall retention of configuration.  相似文献   

17.
Six yeasts were studied for their ability to reduce ethyl 4-chloroacetoacetate (ethyl 4-chloro-3-oxobutanoate) stereoselectively. Five species reduced the substrate to ethyl (S)-4-chloro-3-hydroxybutanoate of high (92-99%) optical purity. With glucose-grown cells, substrate reduction could only be demonstrated when growth was oxygen-limited, whereas xylose-grown Pichia capsulata could be grown under conditions of oxygen excess without losing its reducing ability. Zygosaccha-romyces rouxii exhibited high enantioselectivity (≥98% ee (S)-enantiomer) under all conditions tested, whilst in P. capsulata, a novel switch was observed from producing mainly the (R)-enantiomer using glucose as co-substrate to producing mainly the (R)-enantiomer using 2-propanol as co-substrate. This switch was correlated with a change in reduction predominantly from an NADPH-dependent dehydrogenase system to an NADH-dependent system. In the production of ethyl (R)-4-chloro-3-hydroxybutanoate with P. capsulata, the enantioselectivity was also found to depend upon growth conditions. With glucose-grown cells, higher enantioselectivity was observed using cells harvested in stationary phase (93-94% ee) compared with cells harvested in exponential phase (43-60% ee). Growing P. capsulata with xylose rather than glucose as the major source of carbon for growth resulted in an eight-fold increase in the specific rate of ethyl (R)-4-chloro-3-hydroxybutanoate production using 2-propanol as co-substrate, although enantioselectivity was slightly reduced (65-81% ee) compared with the maximum achieved with glucose-grown cells. The effect of growth on xylose could also be correlated with enhanced activity of an NADH-dependent (R)-selective dehydrogenase system.  相似文献   

18.
Synthesis of the conveniently protected epimer at C-3' of the miharamycin sugar moiety was accomplished starting from the corresponding 3,3'-spiroepoxide. Reaction of the epoxide with lithium cyanide, followed by hydrolysis and spontaneous cyclization, afforded the intermediate deoxylactone methyl 4,6-O-benzylidene-3-C-(carboxymethyl)-alpha-D-glucopyranoside-3',2-lacto ne (8). Stereoselective hydroxylation with MoO5 x py x HMPA, reduction with lithium aluminum hydride and cyclization with diethyl azodicarboxylate-triphenylphosphine gave the target molecule methyl 2,3'-anhydro-4,6-O-benzylidene-3-C-[(R)-1,2-dihydroxyethyl]-alpha -D-glucopyranoside (5). Direct reduction of 8 gave other analogs having no C-3' hydroxyl group together with having a C-3' hydroxyl group (hemiacetal). In addition, C-3' epimers were also synthesized through C-3', C-3' dihydroxy analogs. Wittig reaction of an appropriate ketosugar with [(ethoxycarbonyl)methylene]triphenylphosphorane leading to a 7:3 Z/E mixture, followed by hydroxylation with osmium tetroxide, reduction and cyclization afforded the target molecule 5 and the miharamycin sugar moiety methyl 2,3'-anhydro-4,6-O-benzylidene-3-C-[(S)-1,2-dihydroxyethyl]-alpha -D-glucopyranoside. Examination of X-ray data for 5 and its NMR spectroscopy data allowed us to explain a contradiction reported in the literature.  相似文献   

19.
Islam MT  Tahara S 《Phytochemistry》2000,54(8):901-907
The dihydroflavonols, (2R,3S)-(+)-3',5-dihydroxy-4',7-dimethoxydihydroflavonol and (2R,3R)-(+)-4',5,7-trimethoxydihydroflavonol were isolated from the stem bark of Lannea coromandelica, along with the known (2R,3R)-(+)-4',7-di-O-methyldihydroquercetin, (2R,3R)-(+)-4',7-di-O-methyldihydrokaempferol and (2R,3R)-(+)-4'-O-methyldihydroquercetin. All five compounds were isolated for the first time from the genus Lannea; furthermore, (2R,3S)-(+)-3',5-dihydroxy-4',7-dimethoxydihydroflavonol, was a rare cis-type isomer. The structures of all compounds were elucidated by spectroscopic methods including 2D NMR and CD analysis.  相似文献   

20.
The biotransformation of 1-indanone and 2-indanone to hydroxyindanones was examined with bacterial strains expressing naphthalene dioxygenase (NDO) and toluene dioxygenase (TDO) as well as with purified enzyme components. Pseudomonas sp. strain 9816/11 cells, expressing NDO, oxidized 1-indanone to a mixture of 3-hydroxy-1-indanone (91%) and 2-hydroxy-1-indanone (9%). The (R)-3-hydroxy-1-indanone was formed in 62% enantiomeric excess (ee) (R:S, 81:19), while the 2-hydroxy-1-indanone was racemic. The same cells also formed 2-hydroxy-1-indanone from 2-indanone. Purified NDO components oxidized 1-indanone and 2-indanone to the same products produced by strain 9816/11. P. putida F39/D cells, expressing TDO, oxidized 2-indanone to (S)-2-hydroxy-1-indanone of 76% ee (R:S, 12:88) but did not oxidize 1-indanone efficiently. Purified TDO components also oxidized 2-indanone to (S)-2-hydroxy-1-indanone of 90% ee (R:S, 5:95) and failed to oxidize 1-indanone. Oxidation of 1- and 2-indanone in the presence of [18O]oxygen indicated that the hydroxyindanones were formed by the incorporation of a single atom of molecular oxygen (monooxygenation) rather than by the dioxygenation of enol tautomers of the ketone substrates. As alternatives to chemical synthesis, these biotransformations represent direct routes to 3-hydroxy-1-indanone and 2-hydroxy-1-indanone as the major products from 1-indanone and 2-indanone, respectively.  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