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 171 毫秒
1.
Stereogenic 2-(N-carbamoyl)pyrrolidinylcuprates prepared from scalemic (i.e., enantioenriched) N-Boc-2-lithiopyrrolidine and THF soluble CuCN.2LiCl react with vinyl iodides, vinyl triflates, beta-iodo-alpha,beta-enoates, propargyl mesylates, and allyl bromide to afford the substitution products with excellent enantioselectivity. Excellent enantiomeric ratios are obtained in the conjugate addition reactions with methyl vinyl ketone while low enantiomeric ratios can be achieved with acrylate esters using HMPA/TMSCl activation. Enantiomeric ratios vary with substrate substitution patterns and the observed enantioselectivities appear to be more a function of cuprate-electrophile reactivities than of the reaction type (e.g., substitution, conjugate addition). Low enantiomeric ratios are obtained with the alpha-(N-carbamoyl)benzylcuprates. The lithium-copper transmetalation and cuprate vinylation reactions proceed with retention of configuration.  相似文献   

2.
The conjugate adducts obtained via coupling of alpha-(N-carbamoyl)alkylcuprates with alpha,beta-ynoates, alpha-allenyl esters, or alpha.beta-enoates or enimides undergo N-Boc deprotection and cyclization onto the ester functionality upon treatment with PhOH/TMSCl, catecholboron bromide, or trimethylsilyl triflate. This two-pot sequence provides synthetic routes to 4-alkylidinepyrrolidine-2-ones, 4-alkylidinepyrrolizidin-2-ones, and 4-alkylidineindolizidin-2-ones via allenyl esters; pyrrolin-2-ones, tetrahydropyrrolizin-2-ones, and tetrahydroindolizin-2-ones via alpha/beta-ynoates; pyrrolidin-2-ones, pyrrolizidin-2-ones, and indolizidin-2-ones via alpha,beta-enoates or alpha.beta-enimides. The reluctance of gamma-carbamoyl-alpha,beta-enoates to undergo E/Z isomerization requires the use of (Z)-beta-iodo-alpha,beta-enoates readily prepared by the addition of HI to the alkynyl esters for the efficient preparation of pyrrolinones, tetrahydropyrrolizinones, and tetrahydroindolizinones. Utilization of omega-functionalized alpha,eta-ynoates or beta-iodo-alpha,beta-enoates allows for cyclization onto the omega-functionality providing for a synthetic route to quinolizidines.  相似文献   

3.
Reported here are details on the discovery of a phosphine-catalyzed isomerization of hemiketals and subsequent reactions of the cyclic keto enol ether products. The new cycloisomerization complements a previously reported amine-catalyzed process that gave oxepinones from the same hemiketal starting materials. In the absence of functionality (R(2)) on the cyclic keto enol ether, a rapid and facile dimerization occurs, giving spiroketal products. When the enone is substituted (i.e., R(2) = Ph), the cyclic keto enol ether is sufficiently stable so that it can be isolated; it can then be further reacted in the same pot to provide the corresponding dihydropyrazoles. Both the spiroketal and dihydropyrazole products arise by a tandem reaction that begins with the novel cycloisomerization. The method allows for the rapid introduction of complexity in the products from relatively simple starting materials. It should find application in the synthesis of natural product-like molecules.  相似文献   

4.
The infrared spectra of 3-, 4-, 5- and 6-monochloro-, 3,5,6-trichloro-2-pyridinols (pyridones), and of the unsubstituted 2-pyridinol isolated in a low temperature (20°K) argon matrix were studied. The results indicate that the enol/keto ratio for the isolated pyridinols decreases according to the position of the chlorine atom(s) on the pyridine ring; i.e., 3,5,6-trichloro >-6 > 5 > 4 > 3 monochloro-2-pyridinols. 3,5,6-Trichloro-2-pyridinol exists predominantly in the enol form.  相似文献   

5.
The E- and Z-silyl enol ethers 4 derived from allyl 3-R-3-dimethyl(phenyl)silylpropanoate (R = Me, Pr(i) and Ph) and the Z-silyl enol ethers 7 derived from 4-R-4-dimethyl(phenyl)silylbut-2-enyl acetate (R = Me and Pr(i)) undergo Ireland-Claisen rearrangements largely in the same stereochemical sense, with C-C bond formation taking place anti to the silyl group in the conformations 22, 23 and 24 in which the hydrogen atom on the stereogenic centre is inside, more or less eclipsing the double bond. The E-silyl enol ether E-7a derived from 4-methyl-4-dimethyl(phenyl)silylbut-2-enyl acetate shows low diastereoselectivity in the alternative sense, probably because C-C bond formation takes place anti to the silyl group in the conformation 26 with the methyl group inside, but the silyl enol ether E-7b derived from 4-isopropyl-4-dimethyl(phenyl)silylbut-2-enyl acetate shows low diastereoselectivity in the normal sense. The E- and Z-silyl enol ethers 33 derived from cis-crotyl 3-phenyl-3-dimethyl(phenyl)silylpropanoate and the E-silyl enol ether 39 derived from trans-crotyl 3-phenyl-3-dimethyl(phenyl)silylpropanoate undergo Ireland-Claisen rearrangements largely in the same stereochemical sense as their allyl counterparts, but with moderately high levels of diastereocontrol in setting up the third stereogenic centre following from chair-like transition structures.  相似文献   

6.
The synthetic procedure presented for the preparation of the title compounds requires 1,4-addition of bulky cuprates to cyclohexenones and subsequent reaction with electrophiles. However, the enolates generated by BF(3).OEt(2)-assistance suffer from lack of nucleophilicity. To circumvent this problem, we developed an indirect method consisting of the following three steps: (1) iodination of the cyclohexenones at the alpha position; (2) BF(3).OEt(2)-assisted 1,4-addition of cuprates (Ar(2)Cu(CN)Li(2), Ar = aryl) followed by quenching the enolates with water; (3) reaction of the alpha-iodo-beta-aryl-cylohexanones thus formed with EtMgBr to generate magnesium enolates. The enolates thus generated in this way showed a high reactivity toward ClP(O)(OEt)(2) to furnish enol phosphates. The aforementioned procedure was also applied to a synthesis of optically active Delta(9)-tetrahydrocannabinol. In addition, a naphthalene analogue of the latter compound was also synthesized in a similar way.  相似文献   

7.
Condensation of organic isothiocyanates with cyanoacetamides gave 24 N- and N'-substituted cyanomonothiocarbonylmalonamides in different tautomeric ratios i.e., amide-thioamides (TMA)R3NHCSCH(CN)CONR1R2 (12), thioamide-enols of amides (E) R3NHCSC(CN)=C(OH)NR1R2 (11)or amide-thioenols (TE) R3NHC(SH)=C(CN)CONR1R2 (13). The equilibrium constants (K(thioenol) =[TE]/[TMA] and K(enol) = [E]/[TMA]) in solution depend on R1, R2, R3 and the solvent. The %(E + TE)for NR1R2 increases in the order NMe2 < NHMe < NH2. The (K(thioenol) + K(enol)) in various solvents follows the order CCl4 > CDCl3 > C6D6 > THF-d8 > (CD3)2CO > CD3CN > DMF-d7 > DMSO-d6. The delta(OH) values are 16.46-17.43 and the delta(SH) values are 3.87-5.26 ppm in non polar solvents, e.g.,CDCl3 and 6.34-6.97 ppm in THF-d8 and CD3CN. An intramolecular O-H...O hydrogen bond leads to the preferred Z-configuration of the enols, and an N-H...O bond stabilizes the thioenols' preferred E-configuration with a non-bonded SH in solution. X-Ray crystallography revealed that systems with high %(E + TE) in solution mostly display the enols 11 in the solid state and systems with lower %(E +TE) in solution display structure 12. The differences in delta(OH), delta(NH), K(enol) and crystallographic data for analogous enol and thioenol systems are compared.  相似文献   

8.
A novel convenient method for the generation of thiocarbonyl ylides from readily accessible starting materials and the first synthetic application of in situ generated ylides in the synthesis of silyl enol and dienol ethers, accompanied by C-C bond formation, is described. Under completely neutral conditions without any catalyst or additive, thermal reactions of S-alpha-silylbenzyl thioesters in sealed tubes at 180 degrees C provided silyl enol and dienol ethers in good to excellent yields with high stereoselectivities. This procedure consists of a multistep reaction in a one-pot process, i.e., 1,4-silatropy of S-alpha-silylbenzyl thioesters to give thiocarbonyl ylides, 1,3-electrocyclization of the ylides to give thiiranes, and the extrusion of sulfur from thiiranes to give silyl enol and dienol ethers.  相似文献   

9.
This study investigates how the various components (method, basis set, and treatment of solvent effects) of a theoretical approach influence the relative energies between keto and enol forms of acetylacetone, which is an important model system to study the solvent effects on chemical equilibria from experiment and theory. The computations show that the most popular density functional theory (DFT) approaches, such as B3LYP overestimate the stability of the enol form with respect to the keto form by ~10 kJ mol?1, whereas the very promising SCS‐MP2 approach is underestimating it. MP2 calculations indicate that in particular the basis set size is crucial. The Dunning Huzinaga double ζ basis (D95z(d,p)) used in previous studies overestimates the stability of the keto form considerably as does the popular split‐valence plus polarization (SVP) basis. Bulk properties of the solvent included by continuum approaches strongly stabilize the keto form, but they are not sufficient to reproduce the reversal in stabilities measured by low‐temperature nuclear magnetic resonance experiments in freonic solvents. Enthalpic and entropic effects further stabilize the keto form, however, the reversal is only obtained if also molecular effects are taken into account. Such molecular effects seem to influence only the energy difference between the keto and the enol forms. Trends arising due to variation in the dielectric constant of the solvent result from bulk properties of the solvent, i.e., are already nicely described by continuum approaches. As such this study delivers a deep insight into the abilities of various approaches to describe solvent effects on chemical equilibria. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

10.
The acid dissociation constants of twelve novel drug precursor N-substituted-6-acylbenzothiazolone derivatives were determined by using the UV-vis spectroscopic technique. The protonation and deprotonation behaviors of the investigated molecules were researched from the super basic to super acid regions (i.e., 8 mol·L(-1) KOH to 98% H(2)SO(4)) including the pH region. It is observed that all of the molecules are protonated in the super acidic region. The calculated relative stability values of possible tautomer structures indicate that the keto form of investigated molecules is favored over the enol form. It was predicted that protonation occurs at the amide (oxo) group found in the keto form.  相似文献   

11.
The reaction of lithium (trimethylsilyl)diazomethane with aldehydes and ketones has been investigated, and it has been found that quenching at low temperature with MeOH followed by addition of Rh2(OAc)4 gave silyl enol ethers in high yields. Quenching with other electrophiles (e.g., deuterium, MeI) gave terminal and substituted silyl enol ethers with complete control over regio- and stereochemistry. The mechanism of this novel process has been mapped out through a combination of deuterium labeling, ReactIR, and isolation of reaction intermediates.  相似文献   

12.
Iida A  Osada J  Nagase R  Misaki T  Tanabe Y 《Organic letters》2007,9(10):1859-1862
A pentafluorophenylammonium triflate (PFPAT) catalyst (5 mol %) successfully promoted C-acylation of enol silyl ethers with acid chloride to produce various beta-diketones (12 examples; 62-92% yield). Similarly, C-acylation of ketene silyl acetals or ketene silyl thioacetals (i.e., crossed Claisen condensation) proceeded smoothly to provide not only alpha-monoalkylated beta-keto (thio)esters but also thermodynamically unfavorable (less accessible) alpha,alpha-dialkylated beta-keto (thio)esters in good to excellent yield (38 examples; 60-92% yield).  相似文献   

13.
The first examples of catalytic enantioselective ring-opening/cross-metathesis (EROCM) reactions that involve enol ethers are reported. Specifically, we demonstrate that catalytic EROCM of several oxa- and azabicycles, cyclobutenes and a cyclopropene with an alkyl- or aryl-substituted enol ether proceed readily in the presence of a stereogenic-at-Mo monopyrrolide-monoaryloxide. In some instances, as little as 0.15 mol % of the catalytically active alkylidene is sufficient to promote complete conversion within 10 min. The desired products are formed in up to 90% yield and >99:1 enantiomeric ratio (er) with the disubstituted enol ether generated in >90% Z selectivity. The enol ether of the enantiomerically enriched products can be easily differentiated from the terminal alkene through a number of functionalization procedures that lead to the formation of useful intermediates for chemical synthesis (e.g., efficient acid hydrolysis to afford the enantiomerically enriched carboxaldehyde). In certain cases, enantioselectivity is strongly dependent on enol ether concentration: larger equivalents of the cross partner leads to the formation of products of high enantiomeric purity (versus near racemic products with one equivalent). The length of reaction time can be critical to product enantiomeric purity; high enantioselectivity in reactions that proceed to >98% conversion in as brief a reaction time as 30 s can be nearly entirely eroded within 30 min. Mechanistic rationale that accounts for the above characteristics of the catalytic process is provided.  相似文献   

14.
A new annulation reaction leading to tetrahydropyrans has been discovered. The reaction of homoallylic enol ethers (e.g., 1) with alpha,beta-unsaturated ketones or esters begins with a Mukaiyama-Michael addition. The intermediate oxocarbenium ion undergoes a rapid 2-oxonia-Cope rearrangement, and the resulting zwitterion collapses to form a tetrahydropyran. The reaction is stereoselective with 3-butene-2-one, but leads to diastereomeric mixtures with ethyl acrylate. More complex enones, such as cyclohexenone, also undergo the reaction to produce fused ring products. The optical activity of the substrates is relayed in the tetrahydropyran products.  相似文献   

15.
Superconductors, classified by materials, embrace at least four broad groups: (i) BCS metals and alloys; (ii) heavy Fermion materials; (iii) high- T c cuprates and (some) organic compounds, and (iv) fullerides. Broadly speaking, in classes (i) and (iv), with (i) possibly embracing the recent discovery of superconductivity in MgB 2 with T c ~ 40 K, electron liquids flow through essentially non-magnetic lattices and the electron-phonon interaction is a key component of the mechanism for Cooper pairing. In classes (ii) and (iii), plus the low- T c material Sr 2 RuO 4 , electron or hole liquids flow through assemblies with magnetic spin fluctuations. The nature of the normal state in class (iii) is not yet universally agreed, both Fermi or Luttinger liquids remaining viable to date, the former, however, with the formation of precursor 2 e Bosons somewhat above T c . Our own studies reveal some common ground between classes (ii) and (iii), involving coherence lengths and effective masses, as well as non- s -wave pairing, though the interactions leading to pairing almost certainly have different physical origins in these two groups. Finally, topological superconductivity is reviewed. It is argued that such a treatment of a topological superfluid could eventually deepen the understanding of the class (iv) fullerides. Resonating valence bond theory, used by Anderson and co-workers as an, of course, approximate strongly-correlated electron technique for high- T c cuprates, can itself be re-written in the form of topological superconductivity, as discussed especially by Wiegmann and collaborators.  相似文献   

16.
The Claisen condensation of 3-phosphorylated propionitriles with diethyl oxalate has been examined. The product of the condensation of 3-(methylethoxyphosphinyl)propionitrile, following hydrolysis, has been shown to be the enol form (Z-isomer) of 4-(methylhydroxyphosphinyl)-3-cyano-2-ketoburyric acid. On crystallization from water, the other isomers (the E-isomer of the enol or the keto-forms) were not isolated. Distillation of the products of silylation of the reaction mixture with hexamethyldisilazane gave both geometric isomers of the enol (Z and E). In aqueous solution, the enol gradually cyclizes to give 2-hydroxy-3-(methylhydroxyphosphinyl)citraconimide.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 9, pp. 2120–2127, September, 1990.  相似文献   

17.
From Li+ well-solvating solvents or complex ligands such as THF, [12]crown-4, amines etc., lithium cuprates R2CuLi(*LiX) crystallise in a solvent-separated ion pair (SSIP) structural type (e.g. 10). In contrast, solvents with little donor qualities for Li+ such as diethyl ether or dimethyl sulfide lead to solid-state structures of the contact ion pair (CIP) type (e.g. 11). 1H,6Li HOESY NMR investigations in solutions of R2CuLi(*LiX) (15, 16) are in agreement with these findings: in THF the SSIP 18 is strongly favoured in the equilibrium with the CIP 17, and in diethyl ether one observes essentially only the CIP 17. Salts LiX (X=CN, Cl, Br, I, SPh) have only a minor effect on the ion pair equilibrium. These structural investigations correspond perfectly with Bertz's logarithmic reactivity profiles (LRPs) of reactions of R2CuLi with enones in diethyl ether and THF: the faster reaction in diethyl ether is due to the predominance of the CIP 17 in this solvent, which is the reacting species; in THF only little CIP 17 is present in a fast equilibrium with the SSIP 18. A kinetic analysis of the LRPs quantifies these findings. Recent quantum-chemical studies are also in agreement with the CIP 17 being the reacting species. Thus a uniform picture of structure and reactivity of lithium cuprates emerges.  相似文献   

18.
In this study, we report the first total synthesis and complete stereostructure of gambieric acid A, a potent antifungal polycyclic ether metabolite, in detail. The A/B‐ring exocyclic enol ether 32 was prepared through a Suzuki–Miyaura coupling of the B‐ring vinyl iodide 18 and the alkylborate 33 and subsequent closure of the A‐ring by using diastereoselective bromoetherification as the key transformation. Suzuki–Miyaura coupling of 32 with acetate‐derived enol phosphate 49 , followed by ring‐closing metathesis of the derived diene, produced the D‐ring. Subsequent closure of the C‐ring through a mixed thioacetalization completed the synthesis of the A/BCD‐ring fragment 8 . The A/BCD‐ and F′GHIJ‐ring fragments (i.e., 8 and 9 ) were assembled through Suzuki–Miyaura coupling. The C25 stereogenic center was elaborated by exploiting the intrinsic conformational property of the seven‐membered F′‐ring. After the oxidative cleavage of the F′‐ring, the E‐ring was formed as a cyclic mixed thioacetal (i.e., 70 ) and then stereoselectively allylated by using glycosylation chemistry. Ring‐closing metathesis of the diene 3 thus obtained closed the F‐ring and completed the polycyclic ether skeleton. Finally, the J‐ring side chain was introduced by using a Julia–Kocienski olefination in the presence of CeCl3 to complete the total synthesis of gambieric acid A ( 1 ), thereby unambiguously establishing its complete stereostructure. The present total synthesis enabled us to evaluate the antifungal and antiproliferative activities of 1 and several synthetic analogues.  相似文献   

19.
Laser flash photolyses of 2-, 3-, and 4-diazoacetylpyridines 8 give the corresponding pyridylketenes 7 formed by Wolff rearrangements, as observed by time-resolved infrared spectroscopy, with ketenyl absorptions at 2127, 2125, and 2128 cm(-1), respectively. Photolysis of 2-, 3-, and 4-8 in CH(3)CN containing n-BuNH(2) results in the formation of two transients in each case, as observed by time-resolved IR and UV spectroscopy. The initial transients are assigned as the ketenes 7, and this is confirmed by IR measurements of the decay of the ketenyl absorbance. The ketenes then form the amide enols 12, whose growth and decay are monitored by UV. Similar photolysis of diazoacetophenone leads to phenylketene (5), which forms the amide enol 17. For 3- and 4-pyridylketenes and for phenylketene, the ratios of rate constants for amination of the ketene and for conversion of the amide enol to the amide are 3.1, 7.7, and 22, respectively, while for the 2-isomer the same ratio is 1.8 x 10(7). The stability of the amide enol from 2-7 is attributed to a strong intramolecular hydrogen bond to the pyridyl nitrogen, and this is supported by the DFT calculated structures of the intermediates, which indicate this enol amide is stabilized by 12.8 kcal/mol relative to the corresponding amide enol from phenylketene. Calculations of the transition states indicate a 10.9 kcal/mol higher barrier for conversion of the 2-pyridyl amide enol to the amide as compared to that from phenylketene.  相似文献   

20.
Twenty-three diarylcarbenium ions and 38 pi-systems (arenes, alkenes, allyl silanes and stannanes, silyl enol ethers, silyl ketene acetals, and enamines) have been defined as basis sets for establishing general reactivity scales for electrophiles and nucleophiles. The rate constants of 209 combinations of these benzhydrylium ions and pi-nucleophiles, 85 of which are first presented in this article, have been subjected to a correlation analysis to determine the electrophilicity parameters E and the nucleophilicity parameters N and s as defined by the equation log k(20 degrees C) = s(N + E) (Mayr, H.; Patz, M. Angew. Chem., Int. Ed. Engl. 1994, 33, 938-957). Though the reactivity scales thus obtained cover more than 16 orders of magnitude, the individual rate constants are reproduced with a standard deviation of a factor of 1.19 (Table 1). It is shown that the reactivity parameters thus derived from the reactions of diarylcarbenium ions with pi-nucleophiles (Figure 3) are also suitable for characterizing the nucleophilic reactivities of alkynes, metal-pi-complexes, and hydride donors (Table 2) and for characterizing the electrophilic reactivities of heterosubstituted and metal-coordinated carbenium ions (Table 3). The reactivity parameters in Figure 3 are, therefore, recommended for the characterization of any new electrophiles and nucleophiles in the reactivity range covered. The linear correlation between the electrophilicity parameters E of benzhydryl cations and the corresponding substituent constants sigma(+) provides Hammett sigma(+) constants for 10 substituents from -1.19 to -2.11, i.e., in a range with only very few previous entries.  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