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.
To complete our panorama in structure–activity relationships (SARs) of sandalwood‐like alcohols derived from analogues of α‐campholenal (= (1R)‐2,2,3‐trimethylcyclopent‐3‐ene‐1‐acetaldehyde), we isomerized the epoxy‐isopropyl‐apopinene (?)‐ 2d to the corresponding unreported α‐campholenal analogue (+)‐ 4d (Scheme 1). Derived from the known 3‐demethyl‐α‐campholenal (+)‐ 4a , we prepared the saturated analogue (+)‐ 5a by hydrogenation, while the heterocyclic aldehyde (+)‐ 5b was obtained via a Bayer‐Villiger reaction from the known methyl ketone (+)‐ 6 . Oxidative hydroboration of the known α‐campholenal acetal (?)‐ 8b allowed, after subsequent oxidation of alcohol (+)‐ 9b to ketone (+)‐ 10 , and appropriate alkyl Grignard reaction, access to the 3,4‐disubstituted analogues (+)‐ 4f,g following dehydration and deprotection. (Scheme 2). Epoxidation of either (+)‐ 4b or its methyl ketone (+)‐ 4h , afforded stereoselectively the trans‐epoxy derivatives 11a,b , while the minor cis‐stereoisomer (+)‐ 12a was isolated by chromatography (trans/cis of the epoxy moiety relative to the C2 or C3 side chain). Alternatively, the corresponding trans‐epoxy alcohol or acetate 13a,b was obtained either by reduction/esterification from trans‐epoxy aldehyde (+)‐ 11a or by stereoselective epoxidation of the α‐campholenol (+)‐ 15a or of its acetate (?)‐ 15b , respectively. Their cis‐analogues were prepared starting from (+)‐ 12a . Either (+)‐ 4h or (?)‐ 11b , was submitted to a Bayer‐Villiger oxidation to afford acetate (?)‐ 16a . Since isomerizations of (?)‐ 16 lead preferentially to β‐campholene isomers, we followed a known procedure for the isomerization of (?)‐epoxyverbenone (?)‐ 2e to the norcampholenal analogue (+)‐ 19a . Reduction and subsequent protection afforded the silyl ether (?)‐ 19c , which was stereoselectively hydroborated under oxidative condition to afford the secondary alcohol (+)‐ 20c . Further oxidation and epimerization furnished the trans‐ketone (?)‐ 17a , a known intermediate of either (+)‐β‐necrodol (= (+)‐(1S,3S)‐2,2,3‐trimethyl‐4‐methylenecyclopentanemethanol; 17c ) or (+)‐(Z)‐lancifolol (= (1S,3R,4Z)‐2,2,3‐trimethyl‐4‐(4‐methylpent‐3‐enylidene)cyclopentanemethanol). Finally, hydrogenation of (+)‐ 4b gave the saturated cis‐aldehyde (+)‐ 21 , readily reduced to its corresponding alcohol (+)‐ 22a . Similarly, hydrogenation of β‐campholenol (= 2,3,3‐trimethylcyclopent‐1‐ene‐1‐ethanol) gave access via the cis‐alcohol rac‐ 23a , to the cis‐aldehyde rac‐ 24 .  相似文献   

2.
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   

3.
The synthesis of 3,4‐didehydroionone isomers 4 , (+)‐ 6 , and (?)‐ 6 and of 3,4‐didehydro‐7,8‐dihydroionone isomers 5 , (+)‐ 7 , and (?)‐ 7 was accomplished starting from commercially available racemic α‐ionone ( 1 ). Their preparation of the racemic forms 4 – 7 was first achieved by mean of a number of chemo‐ and regioselective reactions (Schemes 1 and 2). The enantio‐ and diastereoselective lipase‐mediated kinetic acetylation of 4‐hydroxy‐γ‐ionone ( 10a / 10b ) provided 4‐hydroxy‐γ‐ionone (+)‐ 10a /(±)‐ 10b and (+)‐4‐(acetyloxy)‐γ‐ionone ((+) 12b ) (Scheme 3). The latter compounds were used as starting materials to prepare the 3,4‐didehydro‐γ‐ionones (+)‐ and (?)‐ 6 and the 3,4‐didehydro‐7,8‐dihydro‐γ‐ionones (+)‐ and (?)‐ 7 in enantiomer‐enriched form. The absolute configuration of (+)‐ 12b was determine by chemical correlation with (+)‐(6S)‐γ‐ionone ((+)‐ 3 ) and with (?)‐(6S)‐α‐ionone ((?)‐ 1 ) therefore allowing to assign the (S)‐configuration to (+)‐ 6 and (+)‐ 7 . Olfactory evaluation of the above described 3,4‐didehydroionone isomers shows a significant difference between the enantiomers and regioisomers both in fragrance feature and in detection threshold (Table).  相似文献   

4.
Herein, we describe the first total syntheses of five members of the dimeric nuphar alkaloids: (+)‐6,6′‐dihydroxythiobinupharidine (+)‐ 1 a , (+)‐6‐hydroxythiobinupharidine (+)‐ 1 b , (?)‐6,6′‐dihydroxythionuphlutine (?)‐ 2 a , (?)‐6,6′‐dihydroxyneothiobinupharidine (?)‐ 3 a , and (+)‐6,6′‐dihydroxyneothionuphlutine (+)‐ 4 a . The latter two have not been found in nature. We have also made each of their enantiomers (?)‐ 1 a – b , (+)‐ 2 a , (+)‐ 3 a , and (?)‐ 4 a . The key step in these syntheses was the dimerization of an α‐aminonitrile (a hydrolytically stable surrogate for its corresponding hemiaminal) with chiral Lewis acid complexes. We have also reassigned the literature structures of (+)‐ 1 a – 1 b —for those instances in which the NMR spectra were obtained in CD3OD—to their corresponding CD3O‐adducts. Our efforts provide for the first time apoptosis data for (?)‐ 3 a , (+)‐ 4 a , and all five non‐natural enantiomers prepared. The data indicate high apoptotic activity regardless of the enantiomer or relative stereochemical configuration at C7 and C7′.  相似文献   

5.
A methodology for the formation of fibers/gels and vesicles by molecular assembly and for controlling their properties is presented. Two‐component systems of pentamer (P)‐ 5 and tetramer (M)‐ 4 pseudoenantiomeric ethynylhelicenes with decyloxycarbonyl (D) and 4‐methyl‐2‐(2‐methylpropyl)‐1‐pentyloxycarbonyl (bD) side‐chains have been examined. Distinct aggregates were formed by changing the solvent for the three combinations of (P)‐bD‐ 5 /(M)‐bD‐ 4 , (P)‐D‐ 5 /(M)‐bD‐ 4 , and (P)‐D‐ 5 /(M)‐D‐ 4 . In toluene, (P)‐bD‐ 5 /(M)‐bD‐ 4 , (P)‐D‐ 5 /(M)‐bD‐ 4 , and (P)‐D‐ 5 /(M)‐D‐ 4 all formed gels and fibrous assemblies were observed by AFM. The minimum gel‐forming concentration (MGC) decreased in the order (P)‐bD‐ 5 /(M)‐bD‐ 4 >(P)‐D‐ 5 /(M)‐bD‐ 4 >(P)‐D‐ 5 /(M)‐D‐ 4 . In diethyl ether, vesicular formation was observed by dynamic light scattering (DLS), AFM, and TEM, and the size of the vesicles decreased in the order (P)‐bD‐ 5 /(M)‐bD‐ 4 >(P)‐D ‐ 5 /(M)‐bD‐ 4 >(P)‐D ‐ 5 /(M)‐D ‐ 4 . Both fiber/gel and vesicle formation were accompanied by enhanced CDs and redshifted UV/Vis absorption bands with a change in color to deep yellow. These are novel two‐component oligomeric systems that form assemblies of fibers/gels or vesicles depending on the solvent, and the structures and properties of the assemblies can be fine‐tuned by changing the combination of oligomers. In m‐difluorobenzene, a homogeneous solution was obtained with (P)‐D‐ 5 /(M)‐bD‐ 4 , which again exhibits enhanced CDs and redshifted UV/Vis absorptions. Vapor pressure osmometry analysis showed the formation of a bimolecular heteroaggregate. The study has indicated that pseudoenantiomeric oligomers form hetero‐double‐helices that hierarchically assemble to form fibers/gels and vesicles.  相似文献   

6.
On the 1H NMR timescale, 2,2′‐biindolyls with (R)‐configured (1‐alkoxyprop)‐2‐yl, (1‐hydroxyprop)‐2‐yl, or (1‐siloxyprop)‐2‐yl substituents at C‐1 and C‐1′ are atropisomerically stable at <0 °C and interconvert at >30 °C. A 2,2′‐biindolyl (R,R)‐ 17 a of that kind and achiral (!) brominating reagents gave the atropisomerically stable 3,3′‐dibromobiindolyls (M)‐ and/or (P)‐ 18 a at best atropselectively—because of point‐to‐axial asymmetric inductions—and atropdivergently, exhibiting up to 95 % (M)‐ and as much (P)‐atropselectivity. This route to atropisomerically pure biaryls is novel and should extend to other substrates and/or different functionalizations. The dibromobiindolyls (M)‐ and (P)‐ 18 a furnished the biindolyldiphosphanes (M)‐ and (P)‐ 14 without atropisomerization. These syntheses did not require the resolution of a racemic mixture, which distinguishes them from virtually all biaryldiphosphane syntheses known to date. (M)‐ and (P)‐ 14 acted as ligands in catalytic asymmetric allylations and hydrogenations. Remarkably, the β‐ketoester rac‐ 25 c was hydrogenated trans‐selectively with 98 % ee; this included a dynamic kinetic resolution.  相似文献   

7.
The resolution of (1α,5α,6α)‐6‐bromo‐6‐fluoro‐1‐phenylbicyclo[3.1.0]hexane (rac‐ 5) provided the enantiomerically pure precursors (?)‐ 5 and (+)‐ 5 of 1‐phenyl‐1,2‐cyclohexadiene. On treatment of (?)‐ 5 with methyllithium in the presence of 2,5‐dimethylfuran, the pure (?)‐enantiomer of the [4+2] cycloadduct of 2,5‐dimethylfuran onto 1‐phenyl‐1,2‐cyclohexadiene was obtained exclusively. From this result, it is concluded that pure (M)‐1‐phenyl‐1,2‐cyclohexadiene ((M)‐ 7 ) emerged from (?)‐ 5 and was enantiospecifically intercepted to give the product. In the case of indene as trap for (M)‐ 7 , the (?)‐ and the (+)‐enantiomer of the [2+2] cycloadduct were formed in the ratio of 95:5. Highly surprising, remarkable enantioselectivities were also observed, when (M)‐ 7 was trapped with styrene to furnish two diastereomeric [2+2] cycloadducts. Hence, the achiral conformation of the diradical conceivable as intermediate cannot play a decisive part. The enantioselective generation of (M)‐ and (P)‐ 7 by the β‐elimination route was tested as well. Accordingly, 1‐bromo‐2‐phenylcyclohexene was exposed to the potassium salt of (?)‐menthol in the presence of 2,5‐dimethylfuran, and the enantiomeric [4+2] cycloadducts of the latter onto (M)‐ and (P)‐ 7 were produced in the ratio of 55:45.  相似文献   

8.
Due to using (R)‐ or (S)‐α‐methylbenzylamine as a chiral auxiliary, and low‐temperature regime for reduction of the intermediate ferrocenyl‐mono‐ or 1,1′‐bis‐ketimines, the corresponding secondary mono‐ or 1,1′‐bis‐amines were prepared with high diastereoselectivity. Removal of the α‐methylbenzyl group afforded the optically active primary mono‐ and bis‐ferrocenylethylamines in high yields. The absolute configuration of (R,R)‐ 3a and (S,S)‐ 3b was determined by X‐ray single crystal diffraction.  相似文献   

9.
Excitation‐dependent photoluminescence (PL) is a well‐known property of graphene quantum dots (GQDs). For the development of carbon‐based photofunctional materials, GQDs possessing uniform PL properties are in high demand. A protocol has been established to separate spectroscopically uniform lipophilic GQD‐ 1 a from a mixture of GQD‐ 1 mainly composed of GQD‐ 1 a and GQD‐ 1 b . The mixture of GQD‐ 1 was synthesized through the reaction of p‐methoxybenzylamine with GQD‐ 2 prepared from graphite by common oxidative exfoliation. Size‐exclusion chromatography gave rise to GQD‐ 1 a and GQD‐ 1 b , with diameters of 19.8 and 4.9 nm, respectively. Large GQD‐ 1 a showed that the PL was fairly independent of the excitation wavelengths, whereas the PL of small GQD‐ 1 b was dependent on excitation. The excitation‐dependent nature is most likely to be associated with the structures of sp2 domains on the graphene surfaces. The large sp2‐conjugated surface of GQD‐ 1 a is likely to possess well‐developed and large sp2 domains, the band gaps of which do not significantly vary. The small sp2‐conjugated surface of GQD‐ 1 b produces small sp2‐conjugated domains that generate band gaps differing with domain sizes.  相似文献   

10.
The reactivity of allyl alcohols of the pinane series and of their epoxides in the presence of montmorillonite clay in intra‐ and intermolecular reactions was studied. Mutual transformations of (+)‐trans‐pinocarveol ((+)‐ 2 ) and (?)‐myrtenol ((?)‐ 3a ) were major reactions of these compounds on askanite–bentonite clay (Schemes 1 and 2). However, the two reactions gave different isomerization products, indicating that the reactivity of the starting alcohol (+)‐ 2 or (?)‐ 3a was different from that of the same compound (+)‐ 2 or (?)‐ 3 formed in the course of the reactions. (?)‐cis‐ and (+)‐trans‐Verbenol ((?)‐ 16 and (+)‐ 12 , resp.), as well as (?)‐cis‐verbenol epoxide ((?)‐ 20 ) reacted with both aliphatic and aromatic aldehydes on askanite–bentonite clay giving various heterocyclic compounds (Schemes 4, 5 and 7); the reaction path depended on the structure of both the terpenoid and the aldehyde.  相似文献   

11.
A new total synthesis of the marine macrolide (?)‐zampanolide ( 1 ) and the structurally and stereochemically related non‐natural levorotatory enantiomer of (+)‐dactylolide ( 2 ), that is, ent‐ 2 , has been developed. The synthesis features a high‐yielding, selective intramolecular Horner–Wadsworth–Emmons (HWE) reaction to close the 20‐membered macrolactone ring of 1 and ent‐ 2 . The β‐keto phosphonate/aldehyde precursor for the ring‐closure reaction was obtained by esterification of a ω‐diethylphosphono carboxylic acid fragment and a secondary alcohol fragment incorporating the THP ring that is embedded in the macrocyclic core structure of 1 and ent‐ 2 . THP ring formation was accomplished through a segment coupling Prins‐type cyclization. Employing the same overall strategy, 13‐desmethylene‐ent‐ 2 as well as the monocyclic desTHP derivatives of 1 and ent‐ 2 were prepared. Synthetic 1 inhibited human cancer cell growth in vitro with nM IC50 values, while ent‐ 2 , which lacks the diene‐containing hemiaminal‐linked side chain of 1 , is 25‐ to 260‐fold less active. 13‐Desmethylene‐ent‐ 2 as well as the reduced versions of ent‐ 2 and 13‐desmethylene‐ent‐ 2 all showed similar cellular activity as ent‐ 2 itself. The same activity level was attained by the monocyclic desTHP derivative of 1 . Oxidation of the aldehyde functionality of ent‐ 2 gave a carboxylic acid that was converted into the corresponding N‐hexyl amide. The latter showed only μM antiproliferative activity, thus being several hundred‐fold less potent than 1 .  相似文献   

12.
The chiral compounds (R)‐ and (S)‐1‐benzoyl‐2,3,5,6‐tetrahydro‐3‐methyl‐2‐(1‐methylethyl)pyrimidin‐4(1H)‐one ((R)‐ and (S)‐ 1 ), derived from (R)‐ and (S)‐asparagine, respectively, were used as convenient starting materials for the preparation of the enantiomerically pure α‐alkylated (alkyl=Me, Et, Bn) α,β‐diamino acids (R)‐ and (S)‐ 11 – 13 . The chiral lithium enolates of (R)‐ and (S)‐ 1 were first alkylated, and the resulting diasteroisomeric products 5 – 7 were aminated with ‘di(tert‐butyl) azodicarboxylate’ (DBAD), giving rise to the diastereoisomerically pure (≥98%) compounds 8 – 10 . The target compounds (R)‐ and (S)‐ 11 – 13 could then be obtained in good yields and high purities by a hydrolysis/hydrogenolysis/hydrolysis sequence.  相似文献   

13.
Following its resolution by diastereomeric complexation, 5,5′,6,6′‐tetrahydroxy‐3,3,3′,3′‐tetramethyl‐1,1′‐spirobisindane (TTSBI) was used to synthesize a chiral ladder polymer, (+)‐ PIM‐CN . (+)‐ PIM‐COOH was also synthesized by the acid hydrolysis of (+)‐ PIM‐CN . Following characterization, both (+)‐ PIM‐CN and (+)‐ PIM‐COOH were solvent cast directly into semipermeable membranes and evaluated for their ability to enable the selective permeation of a range of racemates, including mandelic acid (Man), Fmoc‐phenylalanine, 1,1′‐bi‐2‐naphthol (binol), and TTSBI. High ee values were observed for a number of analytes, and both materials exhibited high permeation rates. A selective diffusion–permeation mechanism was consistent with the results obtained with these materials. Their high permeability, processability, and ease of chemical modification offer considerable potential for liquid‐phase membrane separations and related separation applications.  相似文献   

14.
The repeating guest units of poly‐(R)‐ 2 were selectively encapsulated by the self‐assembled capsule poly‐ 1 possessing eight polymer side chains to form the supramolecular graft polymer (poly‐ 1 )n?poly‐(R)‐ 2 . The encapsulation of the guest units was confirmed by 1H NMR spectroscopy and the DOSY technique. The hydrodynamic radius of the graft polymer structure was greatly increased upon the complexation of poly‐ 1 . The supramolecular graft polymer (poly‐ 1 )n?poly‐(R)‐ 2 was stably formed in the 1:1 host–guest ratio, which increased the glass transition temperature by more than 10 °C compared to that of poly‐ 1 . AFM visualized that (poly‐ 1 )n?poly‐(R)‐ 2 formed the networked structure on mica. The (poly‐ 1 )n?poly‐(R)‐ 2 gelled in 1,1,2,2‐tetrachloroethane, which led to fabrication of distinct viscoelastic materials that demonstrated self‐healing behavior in a tensile test.  相似文献   

15.
The peculiar and highly diffusive odor signal of flowering clary‐sage plants (Salvia sclarea L.) was identified to derive from trace amounts of 1‐methoxyhexane‐3‐thiol ( 1 ) by mass‐spectrometry analysis and confirmed by comparison with synthetic racemic thiol (±)‐ 1 . The enantiomers (S)‐ and (R)‐ 1 were prepared by enantioselective synthesis, and the absolute configuration of (S)‐ 1 was fully corroborated by X‐ray‐diffraction analysis of the crystalline thioester (1′S,1S)‐ 2 . Compound (S)‐ 1 is one of the most powerful odorants known, with a detection threshold of 0.04⋅10−3 ng/l air, and is, with its herbaceous‐green, alliaceous, and perspiration profile, key to the fragrance of clary‐sage flowers and of the freshly distilled essential oil. As a consequence of its unique odor, 1 was also suspected to be part of the volatiles of a Ruta species where it was subsequently identified together with its homologue, 1‐methoxyheptane‐3‐thiol ( 3 ), 1‐methoxy‐4‐methylpentane‐3‐thiol ( 4 ), and the known 4‐methoxy‐2‐methylbutane‐2‐thiol ( 5 ). The syntheses of (±)‐ 3 and (±)‐ 4 as well as of the enantiomer (R)‐ 4 are described. In both natural fractions, the ratio (S)‐ 1 /(R)‐ 1 was slightly in favor of the (S)‐enantiomer. Natural 4 has (R)‐configuration.  相似文献   

16.
A broadly applicable route to trans‐2,5‐disubstituted pyrrolidines has been developed. Key steps are an asymmetric iridium‐catalyzed allylic amination, a Suzuki–Miyaura coupling, and an intramolecular aza‐Michael addition. Enantiomeric excesses in the range of 93–99 % ee have been achieved. Total syntheses of the alkaloids (?)‐ 225 C , (+)‐ and (?)‐ 223 H (xenovenine), (+)‐ 223 AB , (+)‐ 195 B , and (+)‐ 223 R have been carried out as applications.  相似文献   

17.
The cross‐aldolization of (−)‐(1S,4R,5R,6R)‐6‐endo‐chloro‐5‐exo‐(phenylseleno)‐7‐oxabicyclo[2.2.1]heptan‐2‐one ((−)‐ 25 ) and of (+)‐(3aR,4aR,7aR,7bS)‐ ((+)‐ 26 ) and (−)‐(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazole‐3‐carbaldehyde ((−)‐ 26 ) was studied for the lithium enolate of (−)‐ 25 and for its trimethylsilyl ether (−)‐ 31 under Mukaiyama's conditions (Scheme 2). Protocols were found for highly diastereoselective condensation giving the four possible aldols (+)‐ 27 (`anti'), (+)‐ 28 (`syn'), 29 (`anti'), and (−)‐ 30 (`syn') resulting from the exclusive exo‐face reaction of the bicyclic lithium enolate of (−)‐ 25 and bicyclic silyl ether (−)‐ 31 . Steric factors can explain the selectivities observed. Aldols (+)‐ 27 , (+)‐ 28 , 29 , and (−)‐ 30 were converted stereoselectively to (+)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aR,4aR,7aR,7bS)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]‐furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((+)‐ 62 ), its epimer at the exocyclic position (+)‐ 70 , (−)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((−)‐ 77 ), and its epimer at the exocyclic position (+)‐ 84 , respectively (Schemes 3 and 5). Compounds (+)‐ 62 , (−)‐ 77 , and (+)‐ 84 were transformed to (1R,2R,3S,7R,8S,9S,9aS)‐1,3,4,6,7,8,9,9a‐octahydro‐8‐[(1R,2R)‐1,2,3‐trihydroxypropyl]‐2H‐quinolizine‐1,2,3,7,9‐pentol ( 21 ), its (1S,2S,3R,7R,8S,9S,9aR) stereoisomer (−)‐ 22 , and to its (1S,2S,3R,7R,8S,9R,9aR) stereoisomer (+)‐ 23 , respectively (Schemes 6 and 7). The polyhydroxylated quinolizidines (−)‐ 22 and (+)‐ 23 adopt `trans‐azadecalin' structures with chair/chair conformations in which H−C(9a) occupies an axial position anti‐periplanar to the amine lone electron pair. Quinolizidines 21 , (−)‐ 22 , and (+)‐ 23 were tested for their inhibitory activities toward 25 commercially available glycohydrolases. Compound 21 is a weak inhibitor of β‐galactosidase from jack bean, of amyloglucosidase from Aspergillus niger, and of β‐glucosidase from Caldocellum saccharolyticum. Stereoisomers (−)‐ 22 and (+)‐ 23 are weak but more selective inhibitors of β‐galactosidase from jack bean.  相似文献   

18.
Various ligands, such as (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐oxazol‐2‐yl]ethen‐1‐ol ((S)‐ 1a ) and (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐thiazol‐2‐yl]ethen‐1‐ol ((S)‐ 1c ), were investigated as auxiliaries for the asymmetric synthesis of chiral ruthenium(II) complexes. The reaction of these chiral auxiliary ligands with [RuCl2(dmso)4], 2,2′‐bipyridine (bpy, 2.2 equiv), and triethylamine (10 equiv) in DMF/PhCl (1:8) at 140 °C for several hours diastereoselectively provided the complexes Λ‐[Ru(bpy)2{(S)‐ 1a ? H}] (Λ‐(S)‐ 2a , 52 % yield, 56:1 d.r.) and Λ‐[Ru(bpy)2{(S)‐ 1c ? H}] (Λ‐(S)‐ 2c , 48 % yield, >100:1 d.r.) in a single step after purification. Both Λ‐(S)‐ 2a and Λ‐(S)‐ 2c could be converted into Λ‐[Ru(bpy)3](PF6)2 by replacing the bidentate enolato ligands with bpy, under retention of configuration, induced by either NH4PF6 as a weak acid (from Λ‐(S)‐ 2a : 73 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 77 % yield, 22:1 e.r.), TFA as a strong acid (from Λ‐(S)‐ 2a : 72 % yield, 52:1 e.r.; from Λ‐(S)‐ 2c : 85 % yield, 25:1 e.r.), methylation with Meerwein′s salt (from Λ‐(S)‐ 2a : 59 % yield, 46:1 e.r.; from Λ‐(S)‐ 2c : 86 % yield, 37:1 e.r.), ozonolysis (from Λ‐(S)‐ 2a : 56 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 43 % yield, 6.3:1 e.r.), or oxidation with a peroxy acid (from Λ‐(S)‐ 2a : 72 % yield, 45:1 e.r.; from Λ‐(S)‐ 2c : 79 % yield, 8.5:1 e.r.). This study shows that, except for the reaction with NH4PF6, oxazoline‐enolato complex Λ‐(S)‐ 2a provides Λ‐[Ru(bpy)3](PF6)2 with higher enantioselectivities than analogous thiazoline‐enolato complex Λ‐(S)‐ 2c , which might be due to the higher coordinative stability of the thiazoline‐enolato complex, thus requiring more prolonged reaction times. Thus, this study provides attractive new avenues for the asymmetric synthesis of non‐racemic ruthenium(II)‐polypyridyl complexes without the need for using a strong acid or a strong methylating reagent, as has been the case in all previously reported auxiliary methods from our group.  相似文献   

19.
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   

20.
The reaction of 9H‐fluorene‐9‐thione ( 1 ) with the cis‐ and trans‐isomers of dimethyl 1‐(4‐methoxyphenyl)aziridine‐2,3‐dicarboxylate (cis‐ and trans‐ 2 , resp.) in xylene at 110° yielded exclusively the spirocyclic cycloadduct with trans‐ and cis‐configurations, respectively (trans‐ and cis‐ 3 , resp.; Scheme 1). Analogously, less‐reactive thioketones, e.g., thiobenzophenone ( 5 ), and cis‐ 2 reacted stereoselectively to give the corresponding trans‐1,3‐thiazolidine‐2,4‐dicarboxylate (e.g., trans‐ 8 ; Scheme 2). On the other hand, the reaction of 5 and trans‐ 2 proceeded in a nonstereoselective course to provide a mixture of trans‐ and cis‐substituted cycloadducts. This result can be explained by an isomerization of the intermediate azomethine ylide. Dimethyl 1,3‐thiazolidine‐2,2‐dicarboxylates 14 and 15 were formed in the thermal reaction of dimethyl aziridine‐2,2‐dicarboxylate 11 with aromatic thioketones (Scheme 3). On treatment of 14 and 15 with Raney‐Ni in refluxing EtOH, a desulfurization and ring‐contraction led to the formation of azetidine‐2,2‐dicarboxylates 17 and 18 , respectively (Scheme 4).  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