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.
Palladium‐catalyzed cross‐coupling of various aryl iodides with bicyclopropylidene provided isolable (1′‐arylallylidene)cyclopropanes, which reacted with a number of carbonyl compounds in the presence of Eu(fod)3 under high pressure to furnish oxaspiro[2.5]octene derivatives in moderate to good yields (22–69 %). The reactions of the allylidenecyclopropanes with two azo compounds as dienophiles afforded diazaspiro[2.5]octenes in high yields (82 and 99 %) even at ambient pressure. When treated with nitrosobenzene, two of the allylidenecyclopropanes gave the Diels–Alder adducts in up to 83 and 40 % yield. 2,5‐Diiodo‐p‐xylene coupled twice with bicyclopropylidene, and the product underwent a twofold Diels–Alder reaction with nitrosobenzene to produce the bis(spirocyclopropaneoxazine) derivative in 88 % yield. This overall transformation can be brought about in a one‐pot, two‐step operation by addition of the nitrosoarene to the reaction mixture immediately after formation of the allylidenecyclopropanes to furnish various 5‐oxa‐4‐azaspiro[2.5]oct‐7‐ene derivatives in 22–77 % yield. The coupling of methyl bicyclopropylidenecarboxylate with 2,6‐dimethylphenyl iodide produced a mixture of very stable regioisomeric allylidenecyclopropane derivatives in 90 % yield. The reaction of this mixture with N‐phenyltriazolinedione gave a corresponding mixture of the spirocyclopropanated heterobicycles in 61 % yield.  相似文献   

2.
Hetero Diels–Alder reaction of 3‐butyliminomethyl‐2‐aryl‐1H‐indoles (Schiff's base) 1 with p‐benzoquinone 2 affords six novel 5‐butyl‐11a‐aryl‐4a,5,11,11a‐tetrahydro‐11bH‐indolo[3,2‐c]quinoline‐1,4‐diones 3 in good yields. All the reactions proceeded with complete diastereoselectivity giving only one product in each case, which was characterized on the basis of its elemental analyses and spectral data (IR, 1H NMR, and Mass).  相似文献   

3.
A cationic palladium complex, [Pd(PPh3)2(MeCN)2](BF4)2, catalyzed the carbonylation of 2,3‐dien‐1‐ols under mild conditions. The dienols bearing two or more alkyl substituents on the diene part afforded 1,3‐diene‐2‐carboxylic acids successfully in tetrahydrofuran (THF), while those possessing one or no alkyl substituent gave polymers of the products exclusively. The former afforded the corresponding methyl esters in good yields when the reactions were carried out in methanol, while the latter afforded mainly the Diels–Alder reaction products of the resulting esters. An alkylidene group‐substituted π‐allylpalladium species has been presumed to be an intermediate. Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

4.
The unprecedented phospha‐aza‐Diels–Alder reaction between an activated electron‐poor imine and 2H‐phospholes yields 1‐phospha‐2‐azanorbornenes in a highly chemoselective and moderately diastereoselective reaction. The intermediate 2H‐phospholes, which act as dienes, are formed in situ from the corresponding 1H‐phospholes. Theoretical calculations confirm that the phospha‐aza‐Diels–Alder reaction is of normal electron demand. The reactive P?N bond in 1‐phospha‐2‐azanorbornenes can be cleaved by nucleophiles leading to the formation of 2,3‐dihydrophospholes.  相似文献   

5.
An unprecedented catalytic asymmetric inverse‐electron‐demand aza‐Diels–Alder reaction of indoles with in situ formed azoalkenes is reported. A diverse set of [2,3]‐fused indoline heterocycles were achieved in generally good yields (up to 97 %) with high regioselectivity and diastereoselectivity (>20:1 d.r.), and with excellent enantioselectivity (up to 99 % ee).  相似文献   

6.
Sealed tube reactions of the naphthosultine 8 with a series of electron‐deficient dienophiles (fumaronitrile, N‐phenylmaleimide, dimethyl fumarate, and dimethyl acetylenedicarboxylate) in toluene at 180 °C gave corresponding 1:1 cycloadducts 11–14 in various amounts along with rearranged naphthosulfolene 7 in 67–95% yields. The reaction of 1,2,4,5‐tetra(bromomethyl)benzene with Rongalite (sodium form aldehyde sulfoxylate) and tetrabutylammonium bromide in DMF gave benzodisultines 17 and 18 in a combined yield of 56%. Sealed tube reactions of benzodisultines 17 and 18 with a series of dienophiles in xylene at 200 °C gave corresponding 1:1 and 1:2 cycloadducts 20–27 . The results suggested that thermal extrusion of sulfurdioxide from these sultines led to either o‐naphthoquinodimethane 6 (from 8 ) or bis‐o‐quinodimethane 19 (from 17 and 18 ); sub sequent trapping of these reactive intermediates by dienophiles and SO2 gave various 1:1 and 1:2 Diels‐Alder ad ducts in modest to excellent yields.  相似文献   

7.
Catalyzed by molecular iodine at room temperature, under solvent‐free conditions, a two component aza‐Diels‐Alder cyclization of N‐vinyl‐2‐pyrrolidinone with N‐arylimine gave tetrahydroquinoline derivatives in good yields and high stereo‐selectivity. And three components aza‐Diels‐Alder reaction of N‐vinyl‐2‐pyrrolidinone, anilines and indole‐3‐carbaldehydes under the same condition afford only cis‐product in good yields.  相似文献   

8.
The title compound 2 was readily prepared by MCPBA oxidation of the sulfide 1. Thermal desulfonylation of 2 gave the sulfoxide‐substituted diene 3. The Diels‐Alder reactions of 3 with various dienophiles gave the cyclized products 4‐9 in good yields. The regiochemistry was found to be dominated by the phenylsulfinyl group, but could be reversed by the presence of a Lewis acid. Deprotonation of 2 by BuLi, followed by the reaction with alkyl halides, gave the substituted 2‐sulfolenes 10. A synthetic application of 10 was demonstrated by converting 10e to the bicyclic product 11.  相似文献   

9.
We have studied the solvent, temperature, and pressure influences on the reaction rates of cyclic and acyclic N=N bonds in the Diels–Alder and ene reactions. The transfer from N‐phenylmaleimide ( 9 ) to a structural analogue, 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 2 ), is accompanied by the rate increase in five to six orders of magnitude in the Diels–Alder reactions with cyclopentadiene ( 4 ) and 9,10‐dimethylanthracene ( 5 ), whereas the transfer from dimethyl fumarate ( 10 ) to diethyl azodicarboxylate ( 1 ) increases only in one to two orders of magnitude. The ratio of the reaction rate constants ( 2 + 4 )/( 1 + 4 ) is very large (5.2 × 107) and almost the same (5.3 × 107) as in the ene reactions with tetramethylethylene ( 7 ), ( 2 + 7 )/( 1 + 7 ). It has been observed that the N=N bond in reagent 2 has strong electrophilic, and its N–N moiety in the transition state has nucleophilic properties, which results from the analysis of the solvation enthalpy transfer of reagents, activated complex, and adduct in the Diels–Alder reaction of 2 with anthracene 22 .  相似文献   

10.
Hetero‐Diels–Alder reactions of [60]fullerene with α,β‐unsaturated thio‐oxindoles ( 3a , 3b , 3c ), prepared from thio‐oxindole 1 and heteroaromatic aldehydes ( 2a , 2b , 2c ), to generate tetrahydrothiopyrano[2,3‐b ]indole [60]fullerene cycloadducts ( 5a , 5b , 5c ) under thermal or microwave irradiation were described. The yields were improved, and the reaction time was decreased by conducting the reaction under microwave irradiation.  相似文献   

11.
An efficient and short entry to polyfunctionalized linear triquinanes from 2‐methoxyphenols is described by utilizing the following chemistry. The Diels–Alder reactions of masked o‐benzoquinones, derived from 2‐methoxyphenols, with cyclopentadiene afford tricyclo[5.2.2.02,6]undeca‐4,10‐dien‐8‐ones. Photochemical oxa‐di‐π‐methane (ODPM) rearrangements and 1,3‐acyl shifts of the Diels–Alder adducts are investigated. The ODPM‐rearranged products are further converted to linear triquinanes by using an O‐stannyl ketyl fragmentation. Application of this efficient strategy to the total synthesis of (±)‐Δ9(12)‐capnellene was accomplished from 2‐methoxy‐4‐methylphenol in nine steps with 20 % overall yield.  相似文献   

12.
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   

13.
The first catalytic asymmetric inverse‐electron‐demand (IED) oxa‐Diels–Alder reaction of ortho‐quinone methides, generated in situ from ortho‐hydroxybenzyl alcohols, has been established. By selecting 3‐methyl‐2‐vinylindoles as a class of competent dienophiles, this approach provides an efficient strategy to construct an enantioenriched chroman framework with three adjacent stereogenic centers in high yields and excellent stereoselectivities (up to 99 % yield, >95:5 d.r., 99.5:0.5 e.r.). The utilization of ortho‐hydroxybenzyl alcohols as precursors of dienes and 3‐methyl‐2‐vinylindoles as dienophiles, as well as the hydrogen‐bonding activation mode of the substrates met the challenges of a catalytic asymmetric IED oxa‐Diels–Alder reaction.  相似文献   

14.
We report the combination of transition‐metal‐catalyzed diversified cycloisomerization of 1,6‐enynes with chiral Lewis acid promoted asymmetric Diels–Alder reaction to realize asymmetric cycloisomerization/Diels–Alder relay reactions of 1,6‐enynes with electron‐deficient alkenes. A broad spectrum of chiral [5,6]‐bicyclic products could be acquired in high yields (up to 99 %) with excellent diastereoselectivy (>19:1 dr) and enantioselectivity (up to 99 % ee).  相似文献   

15.
CuI‐catalyzed enantioselective nitroso‐Diels–Alder reactions (NDA reactions) of 2‐nitrosopyridine with various dienes are presented. The [CuPF6(MeCN)4]/Walphos‐CF3 catalyst system is best suited to catalyze the NDA reaction of various dienes by using 2‐nitrosopyridine as a dienophile. In most of the cases studied, cycloadducts are obtained in quantitative yield with very good to excellent enantioselectivities. Based on DFT calculations, a model to explain the stereochemical outcome of the NDA reaction is presented. Finally, an efficient short synthesis of (?)‐peracetylated conduramine A‐1 by applying the enantioselective NDA reaction as a key step is described.  相似文献   

16.
The Diels–Alder reaction between 5‐hydroxy‐1,4‐naphtho­quinone and 5,5‐di­methyl‐3‐vinyl‐1,2‐cyclo­hexa­dienyl acetate by endo addition gives the title compound, C22H22O5, in 68% yield. This racemic diastereoisomer has the opposite regiochemistry to ochromycinone analogues produced previously and may allow access to a new type of anticancer‐active saqua­yamycin analogue.  相似文献   

17.
The bulk free‐radical polymerization of 2‐[(N,N‐dialkylamino)methyl]‐1,3‐butadiene with methyl, ethyl, and n‐propyl substituents was studied. The monomers were synthesized via substitution reactions of 2‐bromomethyl‐1,3‐butadiene with the corresponding dialkylamines. For each monomer the effects of the polymerization initiator, initiator concentration, and reaction temperature on the final polymer structure, molecular weight, and glass‐transition temperature (Tg) were examined. Using 2,2′‐azobisisobutyronitrile as the initiator at 75 °C, the resulting polymers displayed a majority of 1,4 microstructures. As the temperature was increased to 100 and 125 °C using t‐butylperacetate and t‐butylhydroperoxide, the percentage of the 3,4 microstructure increased. Differential scanning calorimetry indicated that all of the Tg values were lower than room temperature. The Tg values were higher when the majority of the polymer structure was 1,4 and decreased as the percentage of the 3,4 microstructure increased. The Diels–Alder side products found in the polymer samples were characterized using NMR and gas chromatography‐mass spectrometry methods. The polymerization temperature and initiator concentration were identified as the key factors that influenced the Diels–Alder dimer yield. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4070–4080, 2000  相似文献   

18.
The catalytic asymmetric [3+2] cycloaddition of α‐ketoesters with 2‐nitrovinylindoles and 2‐nitrovinyl‐ pyrroles has been established. This strategy allowed the construction of structurally diverse pyrrolo[1,2‐a]indoles bearing three contiguous stereocenters in generally high yields and good to excellent stereoselectivities (up to 98% yield, > 98 : 2 dr, 99% ee). The efficient synthesis of tetracyclic psychotropic compound analogue via the derivatization of cycloadduct showed the great synthetic potential of this strategy.  相似文献   

19.
A regioselective sulfonyl/sulfinyl migration cycloisomerization cascade of alkyne‐tethered ynamides is developed in the presence of XPhosgold catalyst. This reaction is the first example of a general [1,3]‐sulfonyl migration from the nitrogen center to the β‐carbon atom of ynamides, followed by umpolung 5‐endo‐dig cyclization of the ynamide α‐carbon atom to the gold‐activated alkyne, and final deaurative [1,5]‐sulfinylation. This process allows the synthesis of peripherally decorated unconventional 4‐sulfinylated pyrroles with broad scope from N‐propargyl‐tethered ynamides. In contrast, N‐homopropargyl‐tethered ynamides undergo intramolecular tetradehydro Diels–Alder reaction to provide 2,3‐dihydro‐benzo[f]indole derivatives. Control experiments and density‐functional theory studies were used to study the reaction pathways.  相似文献   

20.
Highly enantioselective Diels–Alder (DA) and inverse‐electron‐demand hetero‐Diels–Alder (HDA) reactions of β,γ‐unsaturated α‐ketoesters with cyclopentadiene catalyzed by chiral N,N′‐dioxide–Cu(OTf)2 (Tf=triflate) complexes have been developed. Quantitative conversion of β,γ‐unsaturated α‐ketoesters and excellent diastereoselectivities (up to 99:1) and enantioselectivities (up to >99 % ee) were observed for a broad range of substrates. Both aromatic and aliphatic β,γ‐unsaturated α‐ketoesters were found to be suitable substrates for the reactions. Moreover, the chemoselectivity of the DA and HDA adducts were improved by regulating the reaction temperature. Good to high chemoselectivity (up to 94 %) of the DA adducts were obtained at room temperature, and moderate chemoselectivity (up to 65 %) of the HDA adducts were achieved at low temperature. The reaction also featured mild reaction conditions, a simple procedure, and remarkably low catalyst loading (0.1–1.5 mol %). A strong positive nonlinear effect was observed.  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