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.
The complexation between two isomers of citral in lemongrass oil and varying types of cyclodextrins (CDs), α-CD, β-CD, and HP-β-CD, were studied by molecular modeling and physicochemical characterization. The results obtained revealed that the most favorable complex formation governing between citrals in lemongrass oil and CDs were found at a 1:2 mole ratio for all CDs. Complex formation between E-citral and CD was more favorable than between Z-citral and CD. The thermal stability of the inclusion complex was observed compared to the citral in the lemongrass oil. The release time course of citral from the inclusion complex was the diffusion control, and it correlated well with Avrami’s equation. The release rate constants of the E- and Z-citral inclusion complexes at 50 °C, 50% RH were observed at 1.32×10?2 h?1 and 1.43×10?2 h?1 respectively.  相似文献   

2.
Inclusion of the two isomers of citral (E-citral and Z-citral), components of lemongrass oil, was investigated within the confines of various cyclodextrin (α-CD, β-CD and γ-CD) host molecules. Aqueous complex formation constants for E-citral with α-CD, β-CD and γ-CD were determined to be 123, 185, and 204 L/mol, respectively, whereas Z-citral exhibited stronger affinities (157, 206, and 253 L/mol, respectively). The binding trend γ-CD > β-CD > α-CD is a reflection of the more favorable geometrical accommodation of the citral isomers with increasing cavity size. Encapsulation of lemongrass oil within CDs was undertaken through shaking citral:CD (1:1, 1.5:1, and 2:1 molar ratio) mixtures followed by spray drying. Maximum citral retention occurred at a 1:1 molar ratio with β-CD and α-CD demonstrating the highest levels of total E-citral and Z-citral retention, respectively. Furthermore, the β-CD complex demonstrated the slowest release rate of all inclusion complex powders.  相似文献   

3.
All-E-(3R,3′R,6′R)-lutein, all-E-(3R,3′R)-zeaxanthin, all-E-(3R,3′S,6′R)-3′-epilutein and some geometrical isomers of the former two dihydroxycarotenoids have been separated from an extract of human plasma by semipreparative high-performance liquid chromatography on a silica-based nitrile-bonded column. In the order of chromatographic elution, the isolated fractions were identified as all-E-lutein, all-E-zeaxanthin, all-E-3′-epilutein, 9Z-lutein, 9′Z-lutein, a mixture of 13Z-lutein and 13′Z-lutein, 9Z-zeaxanthin, 13Z-zeaxanthin and 15Z-zeaxanthin. The structures of all compounds, including the relative configuration at C(3′) and C(6′) of the luteins and the position of the stereomutated double bonds in the geometrical isomers, were unambiguously established by 1H nuclear magnetic resonance spectroscopy. The absolute configuration of the three all-E compounds was derived by circular dichroism and is also assumed to be valid for the geometrical isomers. The ultraviolet—visible absorption and mass spectra of each of the individually isolated compounds were also in agreement with the proposed structures.  相似文献   

4.
Bacteria, yeasts and filamentous fungi were screened for enantio-specific reduction of the α, β-unsaturated carbon bond in citral to produce citronellal. While a traditional aqueous screening system revealed only Zymomonas mobilis as positive, citronellal was produced in an aqueous/organic two liquid phase system by 11 of the 46 tested strains, which demonstrates the relevance of applying two-phase systems to screening strategies. Z. mobilis and Citrobacter freundii formed 1 mM citronellal in 3 h in the presence of a NADPH regenerating system and 20% (v/v) toluene. In comparison to these bacteria, the eukaryotic strains showed at least five-fold lower citral reductase activities. The bacterial strains produced preferentially the (S)-enantiomer of citronellal with e.e. values of >99% for Z. mobilis and 75% for Citrobacter freundii. In contrast the yeasts produced preferentially (R)-citronellal, i.e. Candida rugosa with an enantiomeric excess value of more than 98%. Many strains formed alcoholic by-products, viz. geraniol, nerol and citronellol. For Z. mobilis the production of these alcohols was suppressed in the presence of various organic solvents, e.g. toluene, and further decreased after EDTA addition.  相似文献   

5.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

6.
A 1:1 mixture of (Z, Z)-3,13-octadecadien-1-ol and its (E, Z)-isomer was synthesized by a Wittig reaction whose stereoselectivity was adjusted to give the products in the required ratio. The corresponding acetate mixture is useful as the attractant for Synanthedon hector.  相似文献   

7.
Chiral aromatic alcohols have received much attention due to their widespread use in pharmaceutical industries. In the asymmetric synthesis processes, the excellent performance of alcohol dehydrogenase makes it a good choice for biocatalysts. In this study, a novel and robust medium-chain alcohol dehydrogenase RhADH from Rhodococcus R6 was discovered and used to catalyse the asymmetric reduction of aromatic ketones to chiral aromatic alcohols. The reduction of 2-hydroxyacetophenone (2-HAP) to (R)-(-)-1-phenyl-1,2-ethanediol ((R)-PED) was chosen as a template to evaluate its catalytic activity. A specific activity of 110 U mg−1 and a 99% purity of e.e. was achieved in the presence of NADH. An efficient bienzyme-coupled catalytic system (RhADH and formate dehydrogenase, CpFDH) was established using a two-phase strategy (dibutyl phthalate and buffer), which highly raised the tolerated substrate concentration (60 g l−1). Besides, a broad range of aromatic ketones were enantioselectively reduced to the corresponding chiral alcohols by this enzyme system with highly enantioselectivity. This system is of the potential to be applied at a commercial scale.  相似文献   

8.
Enzyme extracts of the wild type yeast Zygosaccharomyces bisporus were applied for the pyruvate decarboxylase catalysed condensation of pyruvate and (R)-(+)-and (S)-(?)-perillyl aldehyde, (±)-citronellal, neral, geranial or (R)-(?)-myrtenal to form novel α-hydroxy ketones. Best yields were obtained when the transformation medium contained 25% (v/v) of the cosolvent N,N-dimethylformamide. Conversion of (R)-(+)-perillyl aldehyde to (1R)-1-hydroxy-1-[(4’R)-4’-isopropenyl-1-cyclohexen-1-yl]-2-propanone proceeded highly stereospecifically (>99% de), whereas the stereoselectivity was somewhat less in the transformation of (S)-(?)-perillyl aldehyde (58% de) and (R)-(?)-myrtenal (92% de). All of the new compounds imparted characteristic odour impressions as determined by means of GC-olfactometry.  相似文献   

9.
The tropical root weevil Diaprepes abbreviatus (L. 1758) (Coleoptera: Curculionidae) is a polyphagous insect from the Caribbean Islands and an invasive insect in the southern part of the United States where it is pest of citrus crops and ornamental trees. Adults feed upon foliage where aggregation, mating and oviposition take place. Here, the headspace volatiles from Citrus macrophylla Wester (Rutaceae), D. abbreviatus adults feeding on this plant, adults alone and adult feces, were collected by aeration and solid-phase microextraction (SPME) for analysis by gas chromatography-linked mass spectrometry (GC-MS). Electrophysiological responses of weevil antennal receptors to volatile headspace extracts and synthetic analogues were recorded by gas chromatography-linked electroantennographic detection (GC-EAD) and electroantennograms (EAGs). Antennal responses were recorded to the monoterpenes (R)-(−)-linalool, citronellal, nerol, citral, and geraniol; all present in the headspace of C. macrophylla. Antennal responses were also recorded to carvacrol, present in the headspace of adults and adult feces. The green leaf volatiles cis-3-hexen-1-ol and trans-2-hexen-1-ol, produced by other host plants, elicited reliable responses on the D. abbreviatus antenna. When comparing EAGs between (±)-linalool and (R)-(−)-linalool, no significant difference was found; responses to (R)-(+)-citronellal were larger than for (S)-(−)-citronellal. Among the individual compounds and blends tested using an open T-track dual choice olfactometer, only the blend of (±)-linalool, cis-3-hexen-1-ol and carvacrol (source dose 25:25:2.5 μg) elicited significant attraction of females, the same blend was repellent for males. The biologically active compounds found here likely play a role in host finding by D. abbreviatus and other interactions of the insect with its hostplant. Handling editor: Sam Cook  相似文献   

10.
Old yellow enzymes (OYEs) are potential targets of protein engineering for useful biocatalysts because of their excellent asymmetric reductions of enone compounds. Two OYEs from different yeast strains, Candida macedoniensis AKU4588 OYE (CmOYE) and Pichia sp. AKU4542 OYE (PsOYE), have a sequence identity of 46%, but show different substrate preferences; PsOYE shows 3.4-fold and 39-fold higher catalytic activities than CmOYE toward ketoisophorone and (4S)-phorenol, respectively. To gain insights into structural basis of their different substrate preferences, we have solved a crystal structure of PsOYE, and compared its catalytic site structure with that of CmOYE, revealing the catalytic pocket of PsOYE is wider than that of CmOYE due to different positions of Phe246 (PsOYE)/Phe250 (CmOYE) in static Loop 5. This study shows a significance of 3D structural information to explain the different substrate preferences of yeast OYEs which cannot be understood from their amino acid sequences.

Abbreviations: OYE: Old yellow enzymes, CmOYE: Candida macedoniensis AKU4588 OYE, PsOYE: Pichia sp. AKU4542 OYE  相似文献   


11.
The four possible isomers of tetradeca‐4,8‐dien‐1‐yl acetate and corresponding alcohols were synthesized stereoselectively by synthetic routes employing Wittig coupling reaction for the preparation of (Z,E)‐ and (Z,Z)‐isomers, and alkylation of terminal alkynes for the preparation of (E,E)‐ and (E,Z)‐isomers as the key steps. Synthetic products were characterized by 13C‐ and 1H‐NMR spectroscopy as well as mass‐spectrometric methods. All four isomers gave distinctive mass spectra where m/z 81 fragments clearly dominated. Elution order, followed by retention index presented in parenthesis, of tetradeca‐4,8‐dien‐1‐ols was determined as (Z,Z) (2082.1), (Z,E) (2082.8), (E,E) (2083.1), and (E,Z) (2083.2) from unpolar SPB‐1 column, and as (E,E) (2210.2), (Z,E) (2222.1), (E,Z) (2223.4), and (Z,Z) (2224.7) from polar DB‐WAX column. The isomers of tetradeca‐4,8‐dien‐1‐yl acetates eluted in the order of (Z,Z) (2176.1), (Z,E) (2178.4), (E,Z) (2185.9), and (E,E) (2186.4) from SPB‐1, and (Z,E) (2124.3), (E,E) (2157.7), (Z,Z) (2128.9), and (E,Z) (2135.9) from DB‐WAX columns. Field‐screening tests for attractiveness of tetradeca‐4,8‐dien‐1‐yl acetates revealed that (4Z,8E)‐tetradeca‐4,8‐dien‐1‐yl acetate significantly attracted Phyllonorycter coryli and Chrysoesthia drurella males. (4E,8E)‐Tetradeca‐4,8‐dien‐1‐yl acetate was the most efficient attractant for Ph. esperella and Ph. saportella males, and (4E,8Z)‐tetradeca‐4,8‐dien‐1‐yl acetate was attractive to Ph. cerasicolella males.  相似文献   

12.
We have studied oxygenation of fatty acids by cell extract of Pseudomonas aeruginosa 42A2. Oleic acid ((9Z)-18:1) was transformed to (10S)-hydroperoxy-(8E)-octadecenoic acid ((10S)-HPOME) and to (7S,10S)-dihydroxy-(8E)-octadecenoic acid (7,10-DiHOME). Experiments under oxygen-18 showed that 7,10-DiHOME contained oxygen from air and was formed sequentially from (10S)-HPOME by isomerization. (10R)-HPOME was not isomerized. The (10S)-dioxygenase and hydroperoxide isomerase activities co-eluted on ion exchange chromatography and on gel filtration with an apparent molecular size of ∼50 kDa. 16:1n-7, 18:2n-6, and 20:1n-11 were also oxygenated to 7,10-dihydroxy fatty acids, and (8Z)-18:1 was oxygenated to 6,9-dihydroxy-(7E)-octadecenoic acid. A series of fatty acids with the double bond positioned closer to ((6Z)-18:1, (5Z,9Z)-18:2) or more distant from the carboxyl group ((11Z)-, (13Z)-, and (15Z)-18:1) were poor substrates. The oxygenation mechanism was studied with [7S-2H]18:1n-9, [7R-2H]18:2n-6, and [8R-2H]18:2n-6 as substrates. The pro-R hydrogen at C-8 was lost in the biosynthesis of (10S)-HPODE, whereas the pro-S hydrogen was lost and the pro-R hydrogen was retained at C-7 during biosynthesis of the 7,10-dihydroxy metabolites. Analysis of the fatty acid composition of P. aeruginosa revealed relatively large amounts of (9E/Z)-16:1 and (11E/Z)-18:1 and only traces of 18:1n-9. We found that (11Z)-18:1 (vaccenic acid) was transformed to (11S,14S)-dihydroxy-(12E)-octadecenoic acid and to a mixture of 11- and 12-HPOME, possibly due to reverse orientation of (11Z)-18:1 at the active site compared with oleic acid. The reaction mechanism of the hydroperoxide isomerase suggests catalytic similarities to cytochrome P450.  相似文献   

13.
Linoleate diol synthases (LDS) are homologous 8(R)-dioxygenases with hydroperoxide isomerase activities, expressed in fungal pathogens of humanitarian importance. We report for the first time expression and site-directed mutagenesis of LDS. 7,8-LDS of the take-all fungus, expressed in Pichia pastoris, oxygenated 18:2n − 6 to 8(R)-hydroperoxylinoleic acid, which was unexpectedly isomerized to 5,8(R)-dihydroxylinoleic acid (60% 5S) and to 8(R),13-dihydroxyoctadeca-9(E),11(E)-dienoic acid. The latter was likely formed via hydrolysis of an unstable intermediate, 8(R),9(S)-epoxyoctadeca-10(E),12(Z)-dienoic acid. A tyrosyl radical is formed during 7,8-LDS catalysis, and Tyr376 is the sequence homolog to Tyr385 of cyclooxygenase-1. Tyr376Phe retained hydroperoxide isomerase activity but lacked 8(R)-dioxygenase activity. The putative proximal heme ligand His379 and the N-glycosylation site at Asn216 appeared to be critical for 8(R)-dioxygenase activity, as His379Gln and Asn216Gln were inactive. Treatment with α-mannosidase to shorten N- and O-linked mannosides inhibited the hydroperoxide isomerase but not the 8(R)-dioxygenase. Our results suggest that post-translational modifications may influence the oxidation mechanism of 7,8-LDS.  相似文献   

14.
Aldo-keto reductases (AKRs) are important biocatalysts that can be used to synthesize chiral pharmaceutical alcohols. In this study, the catalytic activity and stereoselectivity of a NADPH-dependent AKR from Kluyveromyces dobzhanskii (KdAKR) toward t-butyl 6-chloro (5S)-hydroxy-3-oxohexanoate ((5S)-CHOH) were improved by mutating its residues in the loop regions around the substrate-binding pocket. And the thermostability of KdAKR was improved by a consensus sequence method targeted on the flexible regions. The best mutant M6 (Y28A/L58I/I63L/G223P/Y296W/W297H) exhibited a 67-fold higher catalytic efficiency compared to the wild-type (WT) KdAKR, and improved R-selectivity toward (5S)-CHOH (dep value from 47.6% to >99.5%). Moreover, M6 exhibited a 6.3-fold increase in half-life (t1/2) at 40°C compared to WT. Under the optimal conditions, M6 completely converted 200 g/L (5S)-CHOH to diastereomeric pure t-butyl 6-chloro-(3R, 5S)-dihydroxyhexanoate ((3R, 5S)-CDHH) within 8.0 h, with a space-time yield of 300.7 g/L/day. Our results deepen the understandings of the structure−function relationship of AKRs, providing a certain guidance for the modification of other AKRs.  相似文献   

15.
(±)-(2Z,4E)-α-Ionylideneacetic acid (2) was enantioselectively oxidized to (?)-(l′S)-(2Z,4E)-4′-hydroxy-α-ionylideneacetic acid (3), (+)-(1′R)-(2Z,4E)-4′-oxo-α-ionylideneacetic acid (4) and (+)-abscisic acid (ABA) (1) by Cercospora cruenta IFO 6164, which can produce (+)-ABA and (+)-4′-oxo-α-acid 4. This metabolism was confirmed by the incorporation of radioactivity from (±)-(2-14C)-(2Z,4E)-α-acid 2 into three metabolites. (?)-4′-Hydroxy-α-acid 3 was a diastereoisomeric mixture consisting of major 1′,4′-trance-4′-hydroxy-α-acid 3a and minor 1′,4′-cis-4′-hydroxy-α-acid 3b. These structures, 3a and 3b, were confirmed by 13C-NMR and 1H-NMR analysis. Also, the enantioselectivity of the microbial oxidation was reexamined by using optically pure α-acid (+)-2 and (?)-2, as the substrates.  相似文献   

16.
Perennial ryegrass (Lolium perenne) is often infected with the fungal‐endophyte Neotyphodium lolii. In addition to the ‘wild‐type’ strain (EWT), several ‘selected’ strains of N. lolii are now being marketed as AR1 (EAR1) and AR37 (EAR37). Each of these strains impact positively on L. perenne's resistance against many insects, including the African black beetles (Heteronychus arator). The impact of volatile oils produced specifically by each strain in the endophyte–grass association in enhancing the grass's resistance to insects is still largely unknown. Keeping these in view, we determined the volatile oil profiles produced by L. perenne infected with either EWT or EAR1 or EAR37 and determined the impacts of these volatiles on the host‐selection behaviour of H. arator adults. In the absence of endophyte infection (E), L. perenne produced 18 different volatile oils. In L. perenne EWT, quantities of 2‐ethyl‐1‐hexanol acetate (Rt = 14.5 min), (Z)‐2‐octen‐1‐ol (Rt = 22.2 min), and butylated hydroxyl toluene (Rt = 23.2 min) were 24, 16 and 26%, respectively, greater than L. perenne E. The strains EAR1 and EAR37 affected differently the quantities of the volatile compounds but not their identity. In the four‐choice bioassay, males and females of H. arator were equally attracted to each strain. In Y‐tube olfactometer, compared against E, H. arator adults were less attracted to L. perenne EWT and EAR1. The attractiveness of EAR37 was similar in effect to E to H. arator. The results indicate that each strain of N. lolii alters the profile of volatile oils in L. perenne differently and that alteration can influence H. arator adult‐host selection.  相似文献   

17.
Variation in coat colour genotypes of archaeological cattle samples from Finland was studied by sequencing 69 base pairs of the extension locus (melanocortin 1‐receptor, MC1R) targeting both a transition and a deletion defining the three main alleles, such as dominant black (ED), wild type (E+) and recessive red (e). The 69‐bp MC1R sequence was successfully analysed from 23 ancient (1000–1800 AD) samples. All three main alleles and genotype combinations were detected with allele frequencies of 0.26, 0.17 and 0.57 for ED, E+ and e respectively. Recessive red and dominant black alleles were detected in both sexes. According to the best of our knowledge, this is the first ancient DNA study defining all three main MC1R alleles. Observed MC1R alleles are in agreement with calculated phenotype frequencies from historical sources. The division of ancient Finnish cattle population into modern Finnish breeds with settled colours was dated to the 20th century. From the existing genotyped populations in Europe (43 breeds, = 2360), the closest match to ancient MC1R genotype frequencies was with the Norwegian native multicoloured breeds. In combined published genotype data of ancient (= 147) and genotypes and phenotypes of modern Nordic cattle (= 738), MC1R allele frequencies showed temporal changes similar to neutral mitochondrial DNA and Y‐chromosomal haplotypes analysed earlier. All three markers indicate major change in genotypes in Nordic cattle from the Late Iron Age to the Medieval period followed by slower change through the historical periods until the present.  相似文献   

18.
[2–14C]-(2Z,4E)-γ-Ionylideneethanol and [2–14C]-(2Z,4E)-γ-ionylideneacetic acid were converted by Cercospora cruenta to [2–14C]-(2Z,4E)-1′,4′-dihydroxy-γ-ionylideneacetic acid and [2-14C]-(2Z,4E)-4′-hydroxy-γ-ionylideneacetic acid, which are intermediates of ABA biosynthesis in C. cruenta.  相似文献   

19.
Several compounds having the basic α-ionylideneacetic acid structure were tested in Cercospora rosicola resuspensions. At 100 μm, all the compounds inhibited abscisic acid (ABA) biosynthesis. Time studies with unlabelled and deuterated (2Z,4E)- and (2E,4E)-α-ionylideneacetic acids showed rapid conversions into both (2Z,4E)- and (2E,4E)-4′-keto-α-ionylideneacetic acids as major products. Incorporation of the label into ABA was specific for the 2Z,4E-isomer. Minor products, identified by GC-MS, were (2Z,4E)- and (2E,4E)-4′-hydroxy-α-ionylideneacetic acids and (2Z,4E)-1′-hydroxy-α-ionylideneacetic acid. The conversion to (2Z,4E)-l′-hydroxy-α-ionylideneacetic acid has not been previously reported and was specific for the 2Z,4E-isomer. A time study for the conversion of methyl esters of [2H3]-(2Z,4E)- and [2H3]-(2E,4E)-4′-keto-α-ionylideneacetates showed a slow introduction of the l′-hydroxyl group and specificity for 2Z,4E-isomer. Conversion of the ethyl esters of (2Z,4E)- and (2E,4E)-l′-hydroxy-α-ionylideneacetates into the ethyl esters of both ABA and (2E,4E)-ABA demonstrated that ABA can be formed by oxidation of the 4′-position after the insertion of the 1′-hydroxy group. The ethyl 1′-hydroxy acids were also isomerized to the corresponding ethyl (2Z,4E)- and ethyl (2E,4E)-3′-hydroxy-β-ionylideneacetates. Ethyl (2Z,4E)-1′-hydroxy acid also gave small amounts of ethyl l′,4′-trans-diol of ABA. These results suggest that ABA may be formed through a (2Z,4E)-1′-hydroxy-α-ionylidene-type intermediate in addition to the previously proposed route through (2Z,4E)-4′-keto-α-ionylideneacetic acid.  相似文献   

20.
The epimeric diterpenes (+)‐(1S,3E,7E,11S,12S)‐verticilla‐3,7‐dien‐12‐ol ( 1 ), isolated from Bursera suntui, and (+)‐(1S,3E,7E,11S,12R)‐verticilla‐3,7‐dien‐12‐ol ( 2 ), isolated from Bursera kerberi, gave the same Wagner‐Meerwein rearrangement product (?)‐(1E,4Z,8Z,11S,12R)‐phomacta‐1,(15)4,8‐triene ( 3 ). The Et2O:BF3‐induced transformations evidence that verticillenes and phomactanes, both containing the bicyclo[9.3.1]pentadecane skeleton, are biogenetically related through the verticillen‐12‐yl cation ( A + ), which also is a key intermediate in the biosynthetic pathways to generate antitumor taxanes. Molecular modeling using the Monte Carlo protocol, followed by density functional theory (DFT) geometry optimization employing the hybrid functionals B3LYP and B3PW91, both with the DGDZVP basis set, secured the configuration of 3 as followed from the good agreement between the calculated and experimental vibrational circular dichroism spectra. Similar DFT calculations allowed determining the absolute configuration of (+)‐(1R,4R,5R,8S,9S,11S,12R,15R)‐1,15:4,5:8,9‐triepoxyphomactane ( 9 ), which surprisingly derives from epoxidation of the second minimum energy conformer of 3 .  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