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 15 毫秒
1.
Three stereoselective syntheses and the physicochemical properties of trans,trans‐5‐(4‐ethoxy‐2,3‐difluorophenyl)‐2‐(4‐propylcyclohexyl)tetrahydropyran, which is an important liquid‐crystal compound with a large negative dielectric anisotropy (Δε=?7.3), are described. The key step in the construction of the trans‐2,5‐disubstituted tetrahydropyran ring in the first approach involved a benzylic cation mediated intramolecular olefin cyclization of a 2‐allyloxy‐1‐arylethanol derivative. The second method included the Et2Zn‐induced 1,2‐aryl shift of a bromohydrin obtained from a hetero‐Diels–Alder reaction, followed by stereoselective bromination. The third approach utilized the hetero‐Diels–Alder reaction of trans‐4‐propylcyclohexanecarboxaldehyde and a 2‐aryl‐3‐(trimethylsilyl)oxy‐1,3‐butadiene, followed by stereoselective protonation. From results obtained by using a quantum chemical calculation method, the reason why the target compound shows a large negative Δε value is discussed.  相似文献   

2.
Since the isolation of brevetoxin‐B, a red tide toxin, many bioactive marine natural products featuring synthetically challenging trans‐fused polycyclic ether ring systems have been reported. We have developed SmI2‐induced cyclization of β‐alkoxyacrylate with aldehyde, affording 2,6‐syn‐2,3‐trans‐tetrahydropyran (THP) or 2,7‐syn‐2,3‐trans‐oxepane with complete stereoselection, as a key reaction of efficient iterative and bi‐directional strategies for the construction of these polycyclic ethers. This reaction is also applicable to the synthesis of 3‐, 5‐, and 6‐methyl‐THPs and 3,5‐dimethyl‐THP. The synthesis of 2‐methyl‐ and 2,6‐dimethyl‐THPs was accomplished by means of a unique methyl insertion. Recently, the SmI2‐induced cyclization was extended to similar reactions using β‐alkoxyvinyl sulfone and sulfoxide. Reaction of (E)‐ and (Z)‐β‐alkoxyvinyl sulfone‐aldehyde afforded 2,6‐syn‐2,3‐trans‐ and 2,6‐syn‐2,3‐cis‐ THPs, respectively. Reaction of (E)‐β‐alkoxyvinyl (R)‐ and (S)‐sulfoxides gave 2,6‐anti‐2,3‐cis‐ and 2,6‐syn‐2,3‐trans‐THPs, respectively. Reaction of (Z)‐β‐alkoxyvinyl (R)‐sulfoxides gave 2,6‐syn‐2,3‐cis‐THP and an olefinic product, while that of (Z)‐β‐alkoxyvinyl (S)‐sulfoxide afforded a mixture of many products. These SmI2‐induced cyclizations have been applied to the total syntheses of various natural products, including brevetoxin‐B, mucocin, pyranicin, and pyragonicin. Synthetic studies on gambierol and maitotoxin are also introduced. © 2010 The Japan Chemical Journal Forum and Wiley Periodicals, Inc. Chem Rec 10: 159–172; 2010: Published online in Wiley InterScience ( www.interscience.wiley.com ) DOI 10.1002/tcr.200900027  相似文献   

3.
A syn‐selective aza‐aldol reaction of boron aza‐enolates, generated from N‐sulfonyl‐1,2,3‐triazoles and 9‐BBN‐H, is reported. It provides a sequential one‐pot procedure for the stereoselective construction of 1,3‐amino alcohols, having contiguous stereocenters, starting from terminal alkynes.  相似文献   

4.
The synthesis of a putative biosynthetic precursor of psymberin including a formal synthesis of the natural product is described. The key step towards the densely functionalized tetrahydropyran core was an enantioselective catalytic Mukaiyama aldol reaction using a titanium(IV)–BINOL catalyst system. syn‐Selective reduction followed by ozonolysis led to a rapid assembly of the tetrahydropyran ring. This flexible approach also allows the synthesis of similar fragments of other complex molecules such as bryostatins and pederins. The syn‐selective coupling between the tetrahydropyran and the aromatic aldehyde was achieved using a boron‐mediated aldol reaction which was followed by further transformations to complete the synthesis of the precursor as well as the formal synthesis of the natural product.  相似文献   

5.
A stereoselective potassium hydride/18-crown-6 mediated double tandem etherification-[2,3]-Wittig-anionic oxy-Cope rearrangement was used to prepare a δ,ε-unsaturated aldehyde with syn stereochemistry. Wittig olefination, dihydroxylation and stereoselective iodine induced cyclisation completed the tetrahydropyran skeleton.  相似文献   

6.
An efficient synthetic route to the concave‐shaped, potentially ionophoric syn‐ and anti‐isomers of 5,6,11,12,17,18‐hexahydro‐5,18:6,11:12,17‐triepoxytrinaphthylene ( 4 ) was elaborated. Starting from ‘oxabenzonorbornadiene’ ( 5 ), the stannylated precursor 9 was prepared in three steps, followed by cyclotrimerization catalyzed by copper(I) thiophene‐2‐carboxylate (CuTC) , which afforded 4 in a syn/anti ratio of 5 : 4.  相似文献   

7.
A minimalist active site redesign of the L ‐fuculose‐1‐phosphate aldolase from E. coli FucA was envisaged, to extend its tolerance towards bulky and conformationally restricted N‐Cbz‐amino aldehyde acceptor substrates (Cbz=benzyloxycarbonyl). Various mutants at the active site of the FucA wild type were obtained and screened with seven sterically demanding N‐Cbz‐amino aldehydes including N‐Cbz‐prolinal derivatives. FucA F131A showed an aldol activity of 62 μmol h?1 mg?1 with (R)‐N‐Cbz‐prolinal, whereas no detectable activity was observed with the FucA wild type. For the other substrates, the F131A mutant gave aldol activities from 4 to about 25 times higher than those observed with the FucA wild type. With regard to the stereochemistry of the reactions, the (R)‐amino aldehydes gave exclusively the anti configured aldol adducts whereas their S counterparts gave variable ratios of anti/syn diastereoisomers. Interestingly, the F131A mutant was highly stereoselective both with (R)‐ and with (S)‐N‐Cbz‐prolinal, exclusively producing the anti and syn aldol adducts, respectively. Molecular models suggest that this improved activity towards bulky and more rigid substrates, such as N‐Cbz‐prolinal, could arise from a better fit of the substrate into the hydrophobic pocket created by the F131A mutation, due to an additional π–cation interaction with the residue K205′ and to efficient contact between the substrate and the mechanistically important Y113′ and Y209′ residues. An expedient synthesis of novel polyhydroxylated pyrrolizidines related to the hyacinthacine and alexine types was accomplished through aldol additions of dihydroxyacetone phosphate (DHAP) to hydroxyprolinal derivatives with the hyperactive FucA F131A as catalyst. The iminocyclitols obtained were fully characterised and found to be moderate to weak inhibitors (relative to 1,4‐dideoxy‐1,4‐imino‐L ‐arabinitol (LAB) and 1,4‐dideoxy‐1,4‐imino‐D ‐arabinitol (DAB)) against glycosidases and rat intestinal saccharidases.  相似文献   

8.
This report delineates our efforts towards the synthesis of a stereochemically well‐defined ketone, the C1?C10 fragment of muamvatin, the first example of a 2, 4, 6‐trioxaadamantane ring skeletal polypropionate marine natural product, using two non‐aldol variants. i) The Shimizu reaction, a Pd(0) mediated stereoselective epoxy‐ring opening of alkenyl oxiranes, was employed for the stereoselective installation of methyl groups in syn‐fashion and ii) Bode's protocol, a NHC‐mediated reaction on β‐epoxy aldehydes, was utilized for stereoselective construction of methyl and hydroxyl groups in anti‐fashion.  相似文献   

9.
A concise asymmetric (>99:1 e.r.) total synthesis of (+)‐anti‐ and (?)‐syn‐mefloquine hydrochloride from a common intermediate is described. The key asymmetric transformation is a Sharpless dihydroxylation of an olefin that is accessed in three steps from commercially available materials. The Sharpless‐derived diol is converted into either a trans or cis epoxide, and these are subsequently converted into (+)‐anti‐ and (?)‐syn‐mefloquine, respectively. The synthetic (+)‐anti‐ and (?)‐syn‐mefloquine samples were derivatized with (S)‐(+)‐mandelic acid tert‐butyldimethylsilyl ether, and a crystal structure of each derivative was obtained. These are the first X‐ray structures for mefloquine derivatives that were obtained by coupling to a known chiral, nonracemic compound, and provide definitive confirmation of the absolute stereochemistry of (+)‐anti‐ as well as (?)‐syn‐mefloquine.  相似文献   

10.
The title compound, C20H18ClNOS2, is a thia­zole‐derived thio­hydroxamic acid O‐ester. The value of Z′ is 3 and the asymmetric unit comprises three mol­ecules of identical helicity along the N—O bond. Two of these show an anti and the third a syn arrangement of substituents attached in positions 3 and 4 to the 1,3‐thia­zole nucleus.  相似文献   

11.
Flash‐vacuum pyrolysis of the quadricyclane derivative 5a at 350° afforded the oxabishomocubane 9a , whose structure was confirmed by an INADEQUATE‐NMR experiment. A computational investigation of the mechanism of this unexpected reaction by DFT and CASPT2‐SCF methods indicated that the reaction path of lowest energy involves a quadricyclane/oxaquadricyclane ( 5 / 12 ) isomerization, followed by a well‐established cycloreversion of 12 to the carbonyl ylide 16 , which subsequently undergoes an intramolecular 1,3‐dipolar cycloadditon to 9 . The lowest‐energy path of the thermal isomerization of the syn‐quadricyclane 6c is its conversion to the syn ‐ sesquinorbornatriene 8c . The corresponding anti‐isomer 5c , however, shows the capability of a degenerate quadricyclane/quadricyclane rearrangement.  相似文献   

12.
A series of 12 analogues of the Cer transfer protein (CERT) antagonist HPA‐12 with long aliphatic chains were prepared as their (1R,3S)‐syn and (1R,3R)‐anti stereoisomers from pivotal chiral oxoamino acids. The enantioselective access to these intermediates as well as their ensuing transformation relied on a practical crystallization‐induced asymmetric transformation (CIAT) process. Sonogashira coupling followed by triple bond reduction and thiophene ring hydrodesulfurization (HDS) into the corresponding alkane moieties was then implemented to complete the synthetic routes delivering the targeted HPA‐12 analogues in concise 4‐ to 6‐step reaction sequences. Ten compounds were evaluated regarding their ability to bind to the CERT START domain by using the recently developed time‐resolved FRET‐based homogeneous (HTR‐FRET) binding assay. The introduction of a lipophilic appendage on the phenyl moiety led to an overall 10‐ to 1000‐fold enhancement of the protein binding, with the highest effect being observed for a n‐hexyl residue in the meta position. The importance of the phenyl ring for the activity was indicated by the reduced potency of the 3‐deoxyphytoceramide aliphatic analogues. The 1,3‐syn stereoisomers were systematically more potent than their 1,3‐anti analogues. In silico studies were used to rationalized these trends, leading to a model of protein recognition coherent with the stronger binding of (1R,3S)‐syn HPAs.  相似文献   

13.
Treatment of (Z)‐1,2,3,4‐tetrakis(pinacolatoboryl)but‐2‐ene, prepared from 2,3‐bis(pinacolatoboryl)buta‐1,3‐diene and bis(pinacolato)diboron, with three molar equivalents of aldehyde in toluene at 100 °C gave the 2,3‐bis(alkylidene)alkane‐1,5‐anti‐diol as a single stereoisomer. The reaction is applicable to both aromatic and α‐unbranched aliphatic aldehydes. The 1,5‐anti‐diols were also synthesized by the one‐pot preparation/triple‐aldehyde addition of the tetraborylated butene. Experimental results for the stepwise treatment of the butene with two types of aldehydes suggest that the rate‐determining step of the triple‐aldehyde addition is the third allylation.  相似文献   

14.
The title compounds were prepared by aldol reaction of anisaldehyde and the respective N,N‐dibenzyl glycinates. Deprotection of the nitrogen atom with Pearlman’s catalyst delivered the unprotected β‐hydroxytyrosine esters, which were further N‐protected as N,N‐phthaloyl (Phth) and N‐fluorenylmethylcarbonyloxy (Fmoc) derivatives. The Friedel–Crafts reaction with various arenes was studied employing these alcohols as electrophiles. It turned out that the facial diastereoselectivitiy depends on the nitrogen protecting group and on the ester group. The unprotected substrates (NH2) gave preferentially syn‐products but the anti‐selectivity increased when going from NHFmoc over NPhth to NBn2. If the ester substituent was varied the syn‐preference increased in the order Me <Et <iPr. The reactions were shown to be fully stereoconvergent and proceeded under kinetic product control. A model is suggested to explain the facial diastereoselectivity based on a conformationally locked benzylic cation intermediate. The reactions are preparatively useful for the N‐unprotected isopropyl ester, which gave Friedel–Crafts alkylation products with good syn‐selectivity (anti/syn=21:79 to 7:93), and for the N,N‐dibenzyl‐protected methyl ester, which led preferentially to anti‐products (anti/syn=80:20 to >95:5). Upon acetylation of the latter compound to the respective acetate, Bi(OTf)3‐catalyzed alkylation reactions became possible, in which silyl enol ethers served as nucleophiles. The respective alkylation products were obtained in high yield and with excellent anti‐selectivitiy (anti/syn≥95:5).  相似文献   

15.
Chlorothiolation of terminal alkynes with sulfenyl chlorides yields anti‐adducts without transition‐metal catalysts. In sharp contrast, transition‐metal‐catalyzed chlorothiolation has not been developed to date, possibly because organosulfur compounds can poison catalyst. Herein, the regio‐ and stereoselective palladium‐catalyzed chlorothiolation of terminal alkynes with sulfenyl chlorides is described. syn‐Chlorothiolation offers a complementary synthetic route to chloroalkenyl sulfides. 2‐Chloroalkenyl sulfides can easily be transformed into various sulfur‐containing products, most of which are often found in natural products and pharmaceuticals.  相似文献   

16.
An efficient total synthesis of the cytotoxic spiroketal natural products aculeatin A and B is described. The synthesis of the 1,3,5‐triol moiety with appropriate configuration was accomplished from the commercially available L ‐malic acid. The key steps in this synthesis are the Barbier allylation, LiAlH4/LiI‐mediated syn‐stereoselective 1,3‐asymmetric reduction, and phenyliodine bis(trifluoroacetate) (=[bis(trifluoroacetoxy)iodo]benzene; PIFA) mediated oxidative spirocyclization.  相似文献   

17.
B3LYP/6‐311+G** calculations were performed systematically on 1,2 (syn) and 1,3 (anti) tautomeric forms of oxa‐ and thia‐ core‐modified porphyrin isomers, which resulted in a total of 86 structures. The structural and energetic variation in all the isomers were analyzed. In corrphycene, hemiporphycene and porphycene the Z forms are more stable compared to the corresponding E forms in both the anti and syn oxa‐ and thiaporphyrin isomers. In contrast, in the syn isomeric forms of [3.0.1.0], [3.1.0.0] and [4.0.0.0] oxaporphyrins and in both syn and anti forms of thiaporphyrin isomers, Z forms are less stable. The HOMO and LUMO values are both negative and varied in a narrow zone, indicating no dramatic effect on the position of heteroatom substitution on the redox properties. The effect of geometric constraints due to the alteration of meso‐bridge length and the hetero atom disposition in the porphyrin core on the relative stabilities of the isomers is analyzed. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

18.
Reaction of [M(NO)Cl3(NCMe)2] (M=Mo, W) with (iPr2PCH2CH2)2PPh (etpip) at room temperature afforded the syn/anti‐[M(NO)Cl3(mer‐etpip)] complexes (M=Mo, a ; W, b ; 3 a,b (syn,anti); syn and anti refer to the relative position of Ph(etpip) and NO). Reduction of 3 a,b (syn,anti) produced [M(NO)Cl2(mer‐etpip)] ( 4 a,b (syn)), [M(NO)Cl(NCMe)(mer‐etpip)] ( 5 a,b (syn,anti)), and [M(NO)Cl(η2‐ethylene)(mer‐etpip)] ( 6 a,b (syn,anti)) complexes. The hydrides [M(NO)H(η2‐ethylene)(mer‐etpip)] ( 7 a,b (syn,anti)) were obtained from 6 a,b (syn,anti) using NaHBEt3 (75 °C, THF) or LiBH4 (80 °C, Et3N), respectively. 7 a,b (syn,anti) were probed in olefin hydrogenations in the absence or presence of a hydrosilane/B(C6F5)3 mixture. The 7 a,b (syn,anti)/Et3SiH/B(C6F5)3 co‐catalytic systems were highly active in various olefin hydrogenations (60 bar H2, 140 °C), with maximum TOFs of 5250 h?1 ( 7 a (syn,anti)) and 8200 h?1 ( 7 b (syn,anti)) for 1‐hexene hydrogenation. The Et3SiH/(B(C6F5)3 co‐catalyst is anticipated to generate a [Et3Si]+ cation attaching to the ONO atom. This facilitates NO bending and accelerates catalysis by providing a vacant site. Inverse DKIE effects were observed for the 7 a (syn,anti)/Et3SiH/(B(C6F5)3 (kH/kD=0.55) and the 7 b (syn,anti)/Et3SiH/(B(C6F5)3 (kH/kD=0.65) co‐catalytic mixtures (20 bar H2/D2, 140 °C).  相似文献   

19.
The fast and highly stereoselective Mannich reaction of aldehydes and ketones with the N‐(p‐methoxyphenyl) ethyl glyoxylate imine catalyzed by polystyrene resins functionalized with (2S,4R)‐hydroxyproline is reported. The effect of the nature of the linker connecting proline with the polymeric backbone has been studied, and a 1,2,3‐triazole linker constructed from azidomethyl polystyrene and O‐propargyl hydroxyproline turns out to be optimal for catalytic activity and enantioselectivity. With aldehyde donors, fast reactions leading to complete conversion in 1–3 h are recorded in DMF. With ketone donors, the reactions tend to be slower, but can be efficiently accelerated (six‐membered ring cycloalkanones) by low‐power microwave irradiation. This approach, which greatly facilitates product isolation since the catalyst is removed by simple filtration, has allowed the implementation of the reactions of aldehyde substrates in a continuous‐flow, single‐pass system. In this manner, the continuous synthesis of the enantiomerically and diastereomerically pure adducts (syn/anti>97:3; ee>99 %) has been achieved at room temperature with residence times of 6.0 min. This methodology has allowed for the preparation of up to 7.8 mmol of the desired Mannich adduct through the use of 0.46 mmol of catalytic resin (5.9 mol %), in a greatly simplified experimental protocol that avoids purification steps.  相似文献   

20.
The NdIII coordination compounds [Nd(4‐pytza)3(H2O)2] · 2H2O ( 1 ) and [Nd(4‐pytza)2(H2O)4]Cl · 2H2O ( 2 ) [H4‐pytza = 5‐(4‐pyridyl)tetrazole‐2‐acetic acid] were synthesized by reactions of K4‐pytza and NdCl3 · 6H2O at different pH values. Single crystal X‐ray diffraction analysis reveals that 4‐pytza ligands in 1 in a μ1,3‐COO synsyn or μ1,1,3‐COO bridging mode coordinate to two central NdIII atoms to display a dinuclear unit, which is connected by one of these 4‐pytza ligands acting in end‐to‐end bridging mode to form a 1D ladder‐like chain. Different from 1 , each 4‐pytza in 2 with a μ1,3‐COO synanti bridging mode coordinates to two NdIII atoms to display a 1D zigzag chain. Furthermore, the luminescence properties of 1 and 2 were investigated at room temperature in the solid state.  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