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 562 毫秒
1.
The formal syntheses of (±)‐nephromopsinic acid, (−)‐phaseolinic acid, and the first total synthesis of (−)‐dihydropertusaric acid from (±)‐ and (−)‐7‐oxabicyclo[2.2.1]hept‐5‐en‐2‐one are described. These syntheses take advantage of a previously reported radical rearrangement (1,2‐acyl migration). A remarkable iodide‐mediated cleavage of a bicyclic system, followed by the introduction of the γ‐chains via a mixed Kolbe electrolysis, are the key steps of these syntheses. This approach is general and could be applied for the preparation of all kinds of paraconic acids with excellent control of the stereochemistry.  相似文献   

2.
The total synthesis of the naturally occurring cyanoglucoside (−)‐bauhinin ( 1 ) was achieved starting from the optically pure oxatrinorbornenone 2 in 12 steps and 8% overall yield. The aglycone of (−)‐bauhinin was easily obtained from the optically pure oxatrinorbornenone derivative 6 by a Wittig‐Horner reaction followed by the opening of the oxa bridge. Glycosidation with tetra‐O‐isobutyryl‐D ‐glucosyl bromide 9 as the reagent in the Koenigs‐Knorr reaction afforded glucoside 10 in 58% yield, which, after photoisomerization and deprotection, gave (−)‐bauhinin ( 1 ).  相似文献   

3.
《化学:亚洲杂志》2017,12(12):1309-1313
An asymmetric route to (−)‐α‐lycorane and (−)‐zephyranthine, and a formal total synthesis of (+)‐clivonine were achieved. A pivotal intermediate, which serves as a potent precursor for the divergent syntheses of these natural products, was accessed by a diastereoselective Pd‐catalyzed cinnamylation of an N tert ‐butanesulfinyl imine.  相似文献   

4.
Outlined herein is a novel and scalable synthesis of (−)‐vindorosine based on two key transformations. A highly diastereoselective vinylogous Mannich addition of dioxinone‐derived lithium dienolates with indolyl N ‐tert‐butanesulfinyl imines has been developed. In addition, an intramolecular Heathcock/aza‐Prins cyclization was introduced to construct both the C, and the highly substituted E rings for the synthesis of (−)‐vindorosine and related alkaloids.  相似文献   

5.
Divergent total syntheses of the enmein‐type natural products (−)‐enmein, (−)‐isodocarpin, and (−)‐sculponin R have been achieved in a concise fashion. Key features of the strategy include 1) an efficient early‐stage cage formation to control succeeding diastereoselectivity, 2) a one‐pot acylation/akylation/lactonization to construct the C‐ring and C8 quarternary center, 3) a reductive alkenylation approach to construct the enmain D/E rings and 4) a flexible route to allow divergent syntheses of three natural products.  相似文献   

6.
The (−)‐ and (+)‐β‐irones ((−)‐ and (+)‐ 2 , resp.), contaminated with ca. 7 – 9% of the (+)‐ and (−)‐transα‐isomer, respectively, were obtained from racemic α‐irone via the 2,6‐trans‐epoxide (±)‐ 4 (Scheme 2). Relevant steps in the sequence were the LiAlH4 reduction of the latter, to provide the diastereoisomeric‐4,5‐dihydro‐5‐hydroxy‐transα‐irols (±)‐ 6 and (±)‐ 7 , resolved into the enantiomers by lipase‐PS‐mediated acetylation with vinyl acetate. The enantiomerically pure allylic acetate esters (+)‐ and (−)‐ 8 and (+)‐ and (−)‐ 9 , upon treatment with POCl3/pyridine, were converted to the β‐irol acetate derivatives (+)‐ and (−)‐ 10 , and (+)‐ and (−)‐ 11 , respectively, eventually providing the desired ketones (+)‐ and (−)‐ 2 by base hydrolysis and MnO2 oxidation. The 2,6‐cis‐epoxide (±)‐ 5 provided the 4,5‐dihydro‐4‐hydroxy‐cisα‐irols (±)‐ 13 and (±)‐ 14 in a 3 : 1 mixture with the isomeric 5‐hydroxy derivatives (±)‐ 15 and (±)‐ 16 on hydride treatment (Scheme 1). The POCl3/pyridine treatment of the enantiomerically pure allylic acetate esters, obtained by enzymic resolution of (±)‐ 13 and (±)‐ 14 , provided enantiomerically pure cisα‐irol acetate esters, from which ketones (+)‐ and (−)‐ 22 were prepared (Scheme 4). The same materials were obtained from the (9S) alcohols (+)‐ 13 and (−)‐ 14 , treated first with MnO2, then with POCl3/pyridine (Scheme 4). Conversely, the dehydration with POCl3/pyridine of the enantiomerically pure 2,6‐cis‐5‐hydroxy derivatives obtained from (±)‐ 15 and (±)‐ 16 gave rise to a mixture in which the γ‐irol acetates 25a and 25b and 26a and 26b prevailed over the α‐ and β‐isomers (Scheme 5). The (+)‐ and (−)‐cisγ‐irones ((+)‐ and (−)‐ 3 , resp.) were obtained from the latter mixture by a sequence involving as the key step the photochemical isomerization of the α‐double bond to the γ‐double bond. External panel olfactory evaluation assigned to (+)‐β‐irone ((+)‐ 2 ) and to (−)‐cisγ‐irone ((−)‐ 3 ) the strongest character and the possibility to be used as dry‐down note.  相似文献   

7.
The first enantioselective total synthesis of (−)‐cycloclavine was accomplished in 8 steps and 7.1 % overall yield. Key features include the first catalytic asymmetric cyclopropanation of allene, mediated by the dirhodium catalyst Rh2(S‐TBPTTL)4, and the enone 1,2‐addition of a new TEMPO carbamate methyl carbanion. An intramolecular strain‐promoted Diels–Alder methylenecyclopropane (IMDAMC) reaction provided a pivotal tricyclic enone intermediate with more than 99 % ee after crystallization. The synthesis of (−)‐ 1 was completed by a late‐stage intramolecular Diels–Alder furan (IMDAF) cycloaddition to install the indole.  相似文献   

8.
The successful application of dihydropyrido[1,2‐a]indolone (DHPI) substrates in Pd‐catalyzed asymmetric allylic alkylation chemistry facilitates rapid access to multiple alkaloid frameworks in an enantioselective fashion. Strategic bromination at the indole C3 position greatly improved the allylic alkylation chemistry and enabled a highly efficient Negishi cross‐coupling downstream. The first catalytic enantioselective total synthesis of (−)‐goniomitine, along with divergent formal syntheses of (+)‐aspidospermidine and (−)‐quebrachamine, are reported herein.  相似文献   

9.
The synthesis of enantiomerically pure (+)‐ and (−)‐γ‐ionone 3 is reported. The first step in the synthesis is the diastereoisomeric enrichment of 4‐nitrobenzoate derivatives of racemic γ‐ionol 12 . The enantioselective lipase‐mediated kinetic acetylation of γ‐ionol 13b afforded the acetate 14 and the alcohol 15 , which are suitable precursors of the desired products (−)‐ and (+)‐ 3 , respectively. The olfactory evaluation of the γ‐ionone isomers shows a great difference between the two enantiomers both in fragrance response and in detection threshold. The selective reduction of (−)‐ 3 and (+)‐ 3 to the γ‐dihydroionones (−)‐(R)‐ 16 and (+)‐(S)‐ 17 , respectively, allowed us to assign unambiguously the absolute configuration of the γ‐ionones.  相似文献   

10.
Treatment of (aR)‐[1,1′binaphthalene]‐8,8′‐diol ((−)‐ 1 ) with hexamethylphosphorous triamide afforded the N,N‐dimethylphosphoramidite (−)‐ 3 (Scheme 1). The synthesis of the analogous N,N‐diisopropylphosphoramidite 4 failed, however, and afforded the acyclic phosphonamidate (−)‐ 5 . The application of the cyclic phosphoramidite (−)‐ 3 towards asymmetric catalysis was investigated. The borane reduction of acetophenone ( 6 ) to (R)‐1‐phenylethanol ( 7 ) in the presence of (−)‐ 3 proceeded with 96% ee (Scheme 2). The use of (−)‐ 3 as ligand in several Cu‐catalyzed addition and substitution reactions resulted in enantioselectivities ranging from 0 to 50% (Schemes 3 and 4).  相似文献   

11.
The total syntheses of (−)‐rotenone and (−)‐dalpanol have been achieved by a group‐selective, stereospecific 1,2‐shift of an epoxy alcohol and SNAr cyclizations. Three oxacycles are constructed, thus illustrating a versatile synthetic route to various rotenoids.  相似文献   

12.
Asymmetric syntheses of the following 17‐membered macrocyclic spermine alkaloids are presented: (−)‐(S)‐protoverbine (=(8S)‐8‐phenyl‐1,5,9,13‐tetraazacycloheptadecane‐6‐one; 1 ), (+)‐(S)‐protomethine (=(2S)‐2‐phenyl‐1,5,9,14‐tetraazabicyclo[12.3.1]octadecan‐4‐one; 2 ), (−)‐(S)‐buchnerine (=(8S)‐8‐(4‐methoxyphenyl)‐1,5,9,13‐tetraazacycloheptadecane‐6‐one; 8 ), (+)‐(S)‐verbamethine (=(+)‐(2S)‐9‐[(E)‐phenylprop‐2‐enoyl]‐2‐phenyl‐1,5,9,14‐tetraazabicyclo[12.3.1]octadecan‐4‐one; 4 ), (−)‐(S)‐verbacine (=(−)‐(8S)‐1‐[(E)‐phenylprop‐2‐enoyl]‐8‐phenyl‐1,5,9,13‐tetraazacycloheptadecan‐6‐one; 3 ), (−)‐(S)‐verbasikrine (=(−)‐(8S)‐1‐[(E)‐3‐(4‐methoxyphenyl)prop‐2‐enoyl]‐8‐phenyl‐1,5,9,13‐tetraazacycloheptadecan‐6‐one; 26 ), (−)‐(S)‐isoverbasikrine (=(−)‐(8S)‐1‐[(Z)‐3‐(4‐methoxyphenyl)prop‐2‐enoyl]‐8‐phenyl‐1,5,9,13‐tetraazacycloheptadecan‐6‐one; 25 ), (+)‐(S)‐verbamekrine (=(+)‐(2S)‐9‐[(E)‐3‐(4‐methoxyphenyl)prop‐2‐enoyl]‐2‐phenyl‐1,5,9,14‐tetraazabicyclo[12.3.1]octadecan‐4‐one; 23 ), and (+)‐(S)‐isoverbamekrine (=(+)‐(2S)‐9‐[(Z)‐3‐(4‐methoxyphenyl)prop‐2‐enoyl]‐2‐phenyl‐1,5,9,14‐tetraazabicyclo[12.3.1]octadecan‐4‐one; 24 ). Effective methods for 1H‐NMR determination of the enantiomeric purity in which (S)‐2‐hydroxy‐2‐phenylacetic acid and (S)‐2‐acetoxy‐2‐phenylacetic acid are used as shift reagents for 1, 8 , and related macrocyclic alkaloids are described.  相似文献   

13.
(−)‐(3S)‐3‐(Tosylamino)butano‐4‐lactone ( 1 ) and its derivative ethyl (−)‐(3S)‐4‐iodo‐3‐(tosylamino)butanoate ( 2 ) are presented as easily accessible chiral building blocks for the construction of a range of different macrolactam frameworks important for the synthesis of naturally occurring polyamine alkaloids as well as for establishing a substance library of such compounds, including S‐containing derivatives for biological tests. In addition to that, the absolute configuration of the spermine alkaloid (−)‐(R)‐budmunchiamine A ( 3 ) from Albizia amara was determined by total synthesis according to the new methodology.  相似文献   

14.
One of a number of intriguing new alkaloids isolated from the Leucetta sp. sponge in 2004, spiroleucettadine displayed unique structural features on a restricted scaffold: a trans ‐fused 5,5‐bicyclic ring system together with an amino hemiketal moiety. Attempts to synthesize the initially proposed structure failed, raising questions as to its veracity, and structure revision ensued in 2008; no successful synthetic approach has been reported to date. Herein, we describe the enantiospecific total synthesis of (−)‐spiroleucettadine by a highly efficient biomimetic approach starting from l ‐tyrosine. One of two key hypervalent‐iodine‐mediated oxidation reactions forged the spirocyclic center, and the other enabled the installation of the methylamine side chain in the penultimate step. Our approach provides synthetic access to a new class of spiroannulated natural products and will enable future studies of the structure–biological‐activity relationships of these antibacterial compounds.  相似文献   

15.
The first total synthesis of the 2,3,5‐O ‐(S ,R )‐nonahydroxytriphenoylated (NHTP) C ‐glucosidic ellagitannin (−)‐vescalin was accomplished through a series of transformations mimicking the sequence of events leading to its biogenesis. The key steps of this synthesis encompass a Wittig‐mediated ring opening of a glucopyranosic hemiacetal, a C‐glucosidation event through a phenolic aldol‐type reaction, and a Wynberg–Feringa–Yamada‐type oxidative phenolic coupling, which forged the NHTP unit of (−)‐vescalin.  相似文献   

16.
A process‐scale stereoselective synthesis of nature‐identical (−)‐(S,S)‐7‐hydroxycalamenal (=(−)‐(5S,8S)‐5,6,7,8‐tetrahydro‐3‐hydroxy‐5‐methyl‐8‐(1‐methylethyl)naphthalene‐2‐carbaldehyde; (−)‐ 1a ) in 96% enantiomeric excess (ee) with the aid of chiral Ru complexes has been developed. The key step was the enantioselective hydrogenation of easily accessible 2‐(4‐methoxyphenyl)‐3‐methylbut‐2‐enoic acid ( 10 ) to (+)‐ 11 in a 86% ee (Scheme 5 and Table 1). A substantial increase in optical purity (96% ee) was achieved by induced crystallization of the intermediate (+)‐3,4‐dihydro‐4‐(1‐methylethyl)‐7‐methoxy‐2H‐naphthalen‐1‐one ((+)‐ 3 ). Computational conformation analysis carried out on the analog (−)‐ 9 rationalized the high diastereoselectivity achieved in the catalytic hydrogenation of the CC bond.  相似文献   

17.
Homologation of readily available α‐boryl pyrrolidines with metal carbenoids is especially challenging even when good leaving groups (Cl) are employed. By performing a solvent switch from Et2O to CHCl3, efficient 1,2‐metalate rearrangement of the intermediate boronate occurs with both halide and ester leaving groups. The methodology was used in the total synthesis of the Stemona alkaloid (−)‐stemaphylline in just 11 steps (longest linear sequence), with high stereocontrol (>20:1 d.r.) and 11 % overall yield. The synthesis also features a late‐stage lithiation–borylation reaction with a tertiary amine containing carbenoid.  相似文献   

18.
A divergent total synthesis of three structurally distinct natural products from imine 9 was accomplished through an approach featuring: 1) a Pd‐catalyzed decarboxylative cross‐coupling, and 2) heteroannulation of 9 with bromoacetaldehyde and oxalyl chloride to give tetrahydroindolizine 6 and dioxopyrrole 7 , respectively. The former was converted into (−)‐rhazinilam, while the latter was converted into (−)‐leucomidine B and (+)‐leuconodine F. A substrate‐directed highly diastereoselective reduction of a sterically unbiased double bond by using a homogeneous palladium catalyst was developed. A self‐induced diastereomeric anisochronism (SIDA) phenomenon was observed for leucomidine B.  相似文献   

19.
The novel tricyclic spiroketal alotane‐type sesterterpenoids showed strikingly different biological activities and potency with subtle structural alterations. Asymmetric total syntheses of the tricyclic sesterterpenoids (−)‐alotaketals A–D and (−)‐phorbaketal A were accomplished [29–31 steps from (−)‐malic acid] in a collective way for the first time. The key features of the strategy included 1) a new cascade cyclization of vinyl epoxy δ‐keto‐alcohols to forge the common tricyclic spiroketal intermediate, 2) a late‐stage allylic C−H oxidation, and 3) olefin cross‐metathesis to install the different side chains.  相似文献   

20.
A reversed‐phase high performance liquid chromatography method has been developed and validated for determination and quantitation of the natural sesquiterpene (−)‐α‐bisabolol. Furthermore the application of the method was done by characterization of chitosan milispheres and liposomes entrapping Zanthoxylum tingoassuiba essential oil, which contains appreciable amount of (−)‐α‐bisabolol. A reversed‐phase C18 column and gradient elution was used with the mobile phase composed of (A) acetonitrile–water–phosphoric acid (19:80:1) and (B) acetonitrile. The eluent was pumped at a flow rate of 0.8 mL/min with UV detection at 200 nm. In the range 0.02–0.64 mg/mL the assay showed good linearity (R2 = 0.9999) and specificity for successful identification and quantitation of (−)‐α‐bisabolol in the essential oil without interfering peaks. The method also showed good reproducibility, demonstrating inter‐day and intra‐day precision based on relative standard deviation values (up to 3.03%), accuracy (mean recovery of 100.69% ± 1.05%) and low values of detection and quantitation limits (0.0005 and 0.0016 mg/mL, respectively). The method was also robust for showing a recovery of 98.81% under a change of solvent in standard solutions. The suitability of the method was demonstrated by the successful determination of association efficiency of the (−)‐α‐bisabolol in chitosan milispheres and liposomes. Copyright © 2009 John Wiley & Sons, Ltd.  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