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.
Manganese(III)-based dioxapropellane synthesis using tricarbonyl compounds   总被引:1,自引:0,他引:1  
Kentaro Asahi 《Tetrahedron》2008,64(8):1620-1634
The manganese(III)-induced oxidative cyclization of 3-(2-oxoethyl)piperidine-2,4-diones was conducted in the presence of 1,1-diarylethenes at reflux temperature to produce 3-aza-7,12-dioxatricyclo[4.3.3.01,6]dodec-8-en-2-ones, simply called azadioxa[4.3.3]propellanes, in excellent yields. A similar oxidation of 2-(2-oxoethyl)cycloalkane-1,3-diones gave the corresponding [4.3.3]-, [5.3.3]-, and [6.3.3]-propellanes. The oxidation of 3-oxopropyl-substituted cycloalkane-1,3-diones also afforded the corresponding propellanes along with the 3-oxopropyl-substituted bicyclic intermediates. The bicyclic intermediates were definitely converted into the corresponding propellanes in the presence of a Lewis acid. The structure determination and the reaction pathway were also described.  相似文献   

2.
The reaction of 2,3-dihydro-1H-cyclopent[b]indole 3 and 1,2,3,4-tetrahydrocarbazole 4 with substituted o-benzoquinones yielded [4.3.3]- and [4.4.3]propellanes, respectively. The physical and chemical properties of the propellane compounds were investigated and a mechanism for the formation of the propellane compounds was discussed.  相似文献   

3.
Summary The synthesis of two new [4.3.3]propellanes is described. The structure of the compounds was determined by chemical and spectroscopic methods. Some aspects of the conformation of the propellanes were studied using1H NMR.
  相似文献   

4.
The effect of organolithium reagent (RLi: R=nBu, iPr, sBu, tBu), solvent (diethyl ether, diethyl ether/THF and MTBE), and stoichiometry on the (?)‐sparteine‐mediated silylation of 7,8‐dipropyltetrathia[7]helicene shows that, unusually, substantially more than 0.5 equivalent of RLi (R=iPr, sBu, tBu) and a large excess of (?)‐sparteine (R=nBu, sBu) is often needed to achieve substantial conversions and good ee values. With nBuLi, however, just one equivalent of the organolithium reagent is sufficient to obtain high conversions. Our best results were obtained using the convenient tBuLi/(?)‐sparteine adduct with which the need for a high (?)‐sparteine/RLi ratio can be avoided. Single‐ and double‐kinetic resolution (KR) procedures give enantiopure samples of 2‐trimethylsilyl‐ and 2,13‐di(trimethylsilyl)‐7,8‐dipropyltetrathia[7]helicene and two‐step double‐KR combining (?)‐sparteine/sBuLi and chiral formamides affords the synthetically valuable 2‐formyl‐7,8‐dipropyltetrathia[7]helicene. This is the first use of (?)‐sparteine for the enantioselective lithiation of helicenes and the first report of tBuLi outperforming sBuLi in a (?)‐sparteine‐mediated procedure.  相似文献   

5.
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   

6.
The 2,2′‐methylenebis[furan] ( 1 ) was converted to 1‐{(4R,6S))‐6‐[(2R)‐2,4‐dihydroxybutyl]‐2,2‐dimethyl‐1,3‐dioxan‐4‐yl}‐3‐[(2R,4R)‐tetrahydro‐4,6‐dihydroxy‐2H‐pyran‐2‐yl)propan‐2‐one ((+)‐ 18 ) and its (4S)‐epimer (?)‐ 19 with high stereo‐ and enantioselectivity (Schemes 13). Under acidic methanolysis, (+)‐ 18 yielded a single spiroketal, (3R)‐4‐{(1R,3S,4′R,5R,6′S,7R)‐3′,4′,5′,6′‐tetrahydro‐4′‐hydroxy‐7‐methoxyspiro[2,6‐dioxabicyclo[3.3.1]nonane‐3,2′‐[2H]pyran]‐6′‐yl}butane‐1,3‐diol ((?)‐ 20 ), in which both O‐atoms at the spiro center reside in equatorial positions, this being due to the tricyclic nature of (?)‐ 20 (methyl pyranoside formation). Compound (?)‐ 19 was converted similarly into the (4′S)‐epimeric tricyclic spiroketal (?)‐ 21 that also adopts a similar (3S)‐configuration and conformation. Spiroketals (?)‐ 20 , (?)‐ 21 and analog (?)‐ 23 , i.e., (1R,3S,4′R,5R,6′R)‐3′,4′,5′,6′‐tetrahydro‐6′‐[(2S)‐2‐hydroxybut‐3‐enyl]‐7‐methoxyspiro[2,6‐dioxabicyclo[3.3.1]nonane‐3,2′‐[2H]pyran]‐4′‐ol, derived from (?)‐ 20 , were assayed for their cytotoxicity toward murine P388 lymphocytic leukemia and six human cancer cell lines. Only racemic (±)‐ 21 showed evidence of cancer‐cell‐growth inhibition (P388, ED50: 6.9 μg/ml).  相似文献   

7.
A concise and enantioselective total synthesis of (+)‐jungermatrobrunin A ( 1 ), which features a unique bicyclo[3.2.1]octene ring skeleton with an unprecedented peroxide bridge, was accomplished in 13 steps by making use of a late‐stage visible‐light‐mediated Schenck ene reaction of (?)‐1α,6α‐diacetoxyjungermannenone C ( 2 ). Along the way, a UV‐light‐induced bicyclo[3.2.1]octene ring rearrangement afforded (+)‐12‐hydroxy‐1α,6α‐diacetoxy‐ent‐kaura‐9(11),16‐dien‐15‐one ( 4 ). These divergent photo‐induced skeletal rearrangements support a possible biogenetic relationship between (+)‐ 1 , (?)‐ 2 , and (+)‐ 4 .  相似文献   

8.
A short and concise synthesis of novel, chiral bicyclo[3.1.0]hex‐2‐ene amino acid derivatives 13 and 14 has been developed. The key step is a stereo‐ and regioselective allylic amination of exo‐ and endo‐methyl bicyclo[3.1.0]hex‐2‐ene‐6‐carboxylates 8 and 9 , which were prepared from 7,7‐dichlorobicyclo[3.2.0]hept‐2‐en‐6‐one ( 1 ). These amino acid derivatives are useful building blocks in medicinal chemistry and can be prepared as chiral compounds by using either (+)‐ 1 or (?)‐ 1 as starting material.  相似文献   

9.
A general stereoselective route to functionalized and substituted tricyclo [5.2.2.0(1,5)]undecenones, tricyclo[6.2.2.0(1,6)]dodecenones, and [3.3.3]- and [4.3.3]propellanes from simple aromatic precursors is reported. The methodology involves generation and cycloaddition of annulated cyclohexa-2,4-dienones with various acrylates followed by manipulation of the resulting tricyclic adducts, leading to functionalized tricyclo[5.2.2.0(1,5)]undecenones and tricyclo[6.2.2.0(1,6)]dodecenones endowed with a beta,gamma-enone chromophore. Photochemical reaction of the tricyclic chromophoric systems followed by reductive cleavage provided an efficient entry into propellanes.  相似文献   

10.
Since 6‐isopropyl‐ ( 11 ) and 6‐isobutyl‐2H‐benzo[b][1,4]dioxepin‐3(4H)‐one ( 12 ) instead of the expected marine odor had been reported to possess lily‐of‐the‐valley notes, albeit weaker than benchmark odorants, the influence of a cyclopropyl ring instead of a methyl branching on the olfactory properties was investigated. 6‐Cyclopropyl‐ ( 27 ), 6‐(2′‐methylcyclopropyl)‐ ( 32 ) and 6‐(cyclopropylmethyl)‐2H‐benzo[b][1,4]dioxepin‐3(4H)‐one ( 39 ) were thus synthesized from 2,3‐dimethoxybenzaldehyde ( 22 ) by a synthetic sequence consisting of Wittig methylenation/ethylenation/homologation with (methoxymethyl)triphenylphosphonium chloride, followed by cyclopropanation, demethylation, Williamson etherification with 3‐chloro‐2‐(chloromethyl)prop‐1‐ene, and Katsuki–Sharpless oxidation. The odor thresholds of the target structures 27 , 32 , and 39 , which are all floral‐green lily‐of‐the‐valley odorants, lie in the range of that of Lilial ( 1 ), with the 6‐cyclopropyl derivative 27 being the most potent (th 0.065 ng/l air). Particularly impressive was the close resemblance of the 6‐(cyclopropylmethyl) derivative 39 with Bourgeonal ( 3 ), which was rationalized by a superposition analysis.  相似文献   

11.
Allylic N‐phenyl imides containing 12‐ and 14‐membered rings, such as compounds 3 and 12 , are easily synthesized by ring enlargement from cycloalkanones and phenyl isocyanates. Irradiation of 3 and 12 in EtOH and MeCN, with high‐ and low‐pressure Hg lamps, led, via the photo‐Fries rearrangement, to the same primary products: the orthocyclophane 8 and the paracyclophane 9 from 3 (Scheme 2), and the corresponding compounds 13 and 14 from 12 (Scheme 3). Besides the primary photorearrangement products, secondary products, the aminocyclophanes 10 and 11 , or 15 and 16 , respectively, were also formed. The total yields of the four products were very high when the N‐phenyl imides were irradiated in MeCN with a low‐pressure Hg lamp: 97 and 93%, respectively. If the para‐position in 3 or 12 is blocked by a Me group, the para‐photo‐Fries rearrangement is prevented. In this case, only one primary photoproduct is formed: the corresponding orthocyclophane ( 17 or 23 , resp.). The most remarkable result was observed on irradiation of the 12‐membered N‐(4‐tolyl) imide 5 in MeCN (low‐pressure lamp). It reacted nearly quantitatively to give only two products: 15‐methyl‐1‐aza[12]orthocyclophane‐2,12‐dione (=16‐methyl‐2‐azabicyclo[12.4.0]octadeca‐1(14),15,17‐triene‐3,13‐dione; 17 ) in 80% yield and 17‐amino‐14‐methyl[11]metacyclophane‐1,11‐dione (=17‐amino‐15‐methylbicyclo[11.3.1]heptadeca‐1(17),13,15‐triene‐2,12‐dione; 19 ) in 16% yield (Scheme 5).  相似文献   

12.
We describe a flexible and divergent route to the pyrrolo‐/pyrido[1,2‐j]quinoline frameworks of tricyclic marine alkaloids via a common intermediate formed by the ester–enolate Claisen rearrangement of a cyclic amino acid allylic ester. We have synthesized the proposed structure of polycitorols and demonstrated that the structure of these alkaloids requires revision. In addition to asymmetric formal syntheses, stereoselective and concise total syntheses of (?)‐lepadiformine and (?)‐fasicularin were also accomplished from simple, commercially available starting materials in a completely substrate‐controlled manner. The key step in these total syntheses was the reagent‐dependent stereoselective reductive amination of the common intermediate to yield either indolizidines 55 a or 55 b . Aziridinium‐mediated carbon homologation of the hindered C‐10 group to the homoallylic group facilitated the synthesis.  相似文献   

13.
A number of aryl 3‐arylprop‐2‐ynoates 3 has been prepared (cf. Table 1 and Schemes 3 – 5). In contrast to aryl prop‐2‐ynoates and but‐2‐ynoates, 3‐arylprop‐2‐ynoates 3 (with the exception of 3b ) do not undergo, by flash vacuum pyrolysis (FVP), rearrangement to corresponding cyclohepta[b]furan‐2(2H)‐ones 2 (cf. Schemes 1 and 2). On melting, however, or in solution at temperatures >150°, the compounds 3 are converted stereospecifically to the dimers 3‐[(Z)‐diarylmethylidene]‐2,3‐dihydrofuran‐2‐ones (Z)‐ 11 and the cyclic anhydrides 12 of 1,4‐diarylnaphthalene‐2,3‐dicarboxylic acids, which also represent dimers of 3 , formed by loss of one molecule of the corresponding phenol from the aryloxy part (cf. Scheme 6). Small amounts of diaryl naphthalene‐2,3‐dicarboxylates 13 accompanied the product types (Z)‐ 11 and 12 , when the thermal transformation of 3 was performed in the molten state or at high concentration of 3 in solution (cf. Tables 2 and 4). The structure of the dihydrofuranone (Z)‐ 11c was established by an X‐ray crystal‐structure analysis (Fig. 1). The structures of the dihydrofuranones 11 and the cyclic anhydrides 12 indicate that the 3‐arylprop‐2‐ynoates 3 , on heating, must undergo an aryl O→C(3) migration leading to a reactive intermediate, which attacks a second molecule of 3 , finally under formation of (Z)‐ 11 or 12 . Formation of the diaryl dicarboxylates 13 , on the other hand, are the result of the well‐known thermal Diels‐Alder‐type dimerization of 3 without rearrangement (cf. Scheme 7). At low concentration of 3 in decalin, the decrease of 3 follows up to ca. 20% conversion first‐order kinetics (cf. Table 5), which is in agreement with a monomolecular rearrangement of 3 . Moreover, heating the highly reactive 2,4,6‐trimethylphenyl 3‐(4‐nitrophenyl)prop‐2‐ynonate ( 3f ) in the presence of a twofold molar amount of the much less reactive phenyl 3‐(4‐nitrophenyl)prop‐2‐ynonate ( 3g ) led, beside (Z)‐ 11f , to the cross products (Z)‐ 11fg , and, due to subsequent thermal isomerization, (E)‐ 11fg (cf. Scheme 10), the structures of which indicated that they were composed, as expected, of rearranged 3f and structurally unaltered 3g . Finally, thermal transposition of [17O]‐ 3i with the 17O‐label at the aryloxy group gave (Z)‐ and (E)‐[17O2]‐ 11i with the 17O‐label of rearranged [17O]‐ 3i specifically at the oxo group of the two isomeric dihydrofuranones (cf. Scheme 8), indicating a highly ordered cyclic transition state of the aryl O→C(3) migration (cf. Scheme 9).  相似文献   

14.
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   

15.
The 3‐methyl‐4‐(tricyclo[5.2.1.02,6]dec‐4‐en‐8‐ylidene)butan‐2‐ols (=Fleursandol®; rac‐ 10 ), a new class of sandalwood odorants, were synthesized in their enantiomerically pure forms by use of tricyclo[5.2.1.02,6]dec‐4‐en‐8‐ones 17 and ent‐ 17 and (tetrahydro‐2H‐pyran‐2‐yl)‐protected 4‐bromo‐3‐methylbutan‐2‐ols 22 and ent‐ 22 as starting materials (Schemes 2–4). Only four of 16 possible stereoisomers of rac‐ 10 possess the typical, very pleasant, long‐lasting sandalwood odor (Table 1). The (2S,3R,4E,1′R,2′R,6′R,7′R)‐isomer ent‐ 10a is by far the most important representative, with an odor threshold of 5 μg/l in H2O.  相似文献   

16.
Discrepancies between chiroptical data from the literature and our determination of the structure of the title compounds (+)‐ 5 and (+)‐ 9a were resolved by an unambiguous assignment of their absolute configuration. Accordingly, the dextrorotatory cis‐3‐hydroxy esters have (3R,4R)‐ and the laevorotatory enantiomers (3S,4S)‐configuration. The final evidences were demonstrated on both enantiomers (+)‐ and (?)‐ 5 by biological reduction of 4 by bakers' yeast and stereoselective [RuII(binap)]‐catalyzed hydrogenations of 4 (Scheme 2), by the application of the NMR Mosher method on (+)‐ and (?)‐ 5 (Scheme 3), as well as by the transformation of (+)‐ 5 into a common derivative and chiroptical correlation (Scheme 4).  相似文献   

17.
(?)‐Daphnilongeranin B and (?)‐daphenylline are two hexacyclic Daphniphyllum alkaloids, each containing a complex cagelike backbone. Described herein are the first asymmetric total synthesis of (?)‐daphnilongeranin B and a bioinspired synthesis of (?)‐daphenylline with an unusual E ring embedded in a cagelike framework. The key features include an intermolecular [3+2] cycloaddition, a late‐stage aldol cyclization to install the F ring of daphnilongeranin B, and a bioinspired cationic rearrangement leading to the tetrasubstituted benzene ring of daphenylline.  相似文献   

18.
A concise and efficient route for the synthesis of spiro‐pyranopyrazoles and oxa‐aza‐[3.3.3]propellanes by simple regioselective multicomponent reaction of ninhydrin, malononitrile, hydrazine derivatives, and β‐keto esters or dimethyl acetylenedicarboxylate was developed. This protocol provides an alternative method for combinatorial and parallel syntheses in drug discovery. The value of this method lies in its simplicity, regioselectivity, and good yields. The structures of 3 and 4 were corroborated spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS). A plausible mechanism for this type of reaction is proposed (Schemes 2 and 3).  相似文献   

19.
The first total synthesis of the natural product (?)‐(19R)‐ibogamin‐19‐ol ((?)‐ 1 ) is reported (biogenetic atom numbering). Starting with L ‐glutamic acid from the chiral pool and (2S)‐but‐3‐en‐2‐ol, the crucial aliphatic isoquinuclidine (= 2‐azabicyclo[2.2.2]octane) core containing the entire configurational information of the final target was prepared in 15 steps (overall yield: 15%). The two key steps involved a highly effective, self‐immolating chirality transfer in an Ireland–Claisen rearrangement and an intramolecular nitrone‐olefin 1,3‐dipolar cycloaddition reaction (Scheme 3). Onto this aliphatic core was grafted the aromatic moiety in the form of N(1)‐protected 1H‐indole‐3‐acetic acid by application of the dicyclohexylcarbodiimide (DCC) method (Scheme 4). Four additional steps were required to adjust the substitution pattern at C(16) and to deprotect the indole subunit for the closure of the crucial 7‐membered ring present in the targeted alkaloid family (Schemes 4 and 5). The spectral and chiroptical properties of the final product (?)‐ 1 matched the ones reported for the naturally occurring alkaloid, which had been isolated from Tabernaemonatana quadrangularis in 1980. The overall yield of the entire synthesis involving a linear string of 20 steps amounted to 1.9% (average yield per step: 82%).  相似文献   

20.
Starting from inexpensive (E)‐β‐farnesene ( 1 ), an eight‐step enantioselective synthesis of the olfactively precious Ambrox® ((?)‐ 2a ) has been performed. The crucial step is the catalytic asymmetric isomerization of (2E,6E)‐N,N‐diethylfarnesylamine ( 3 ) to the corresponding enamine (?)‐(R,E)‐ 4a , applying Takasago's well‐known industrial methodology. The resulting dihydrofarnesal ((+)‐(R)‐ 5 ) (90% yield, 96% ee), obtained after in situ hydrolysis (AcOH, H2O), was then cyclized under catalytic SnCl4 conditions, via its corresponding unreported enol acetate (?)‐(R)‐ 4b , to afford trans‐decalenic aldehyde (+)‐ 6a . Subsequent transformations furnished bicyclic ketone (?)‐ 8a and unsaturated nitrile (+)‐ 11 , both reported as intermediates to access to (?)‐ 2a .  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