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 625 毫秒
1.
We present here a novel programmable polymerization route for the synthesis of new indole‐based polymers via a catalyst‐free nucleophilic substitution reaction. The polycondensation of 4‐hydroxyindole with different activated difluoro monomers undergoes in N‐methylpyrrolidinone, affording soluble poly(N‐aryleneindole ether)s (PEINs) with high molecular weights (Mw up to 486,000) in high yields (up to 96%). The structures of the polymers are characterized by means of FT‐IR, 1H NMR spectroscopy and elemental analysis, the results show good agreement with the proposed structures. The resulting polymers are processable and enjoy high glass transition temperatures (Tgs > 180 °C) and thermal stability (Tds > 420 °C). Thin films of PEINs show great mechanical behaviors with high tensile strength up to 104 Mpa, and good optical transparency. In addition, due to the indole moieties in the main chains, all these PEINs are endowed with significantly strong photonic luminescence in chloroform and display highly solvent‐dependent emission bands. Especially, the polymer PEIN‐3 carrying sulfonyl units, shows outstanding blue‐light emission with high quantum yields (45.2%, determined against quinine sulfate). The results obtained by cyclic voltammetry suggest that PEINs possess good electroactivity. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 313–320  相似文献   

2.
The bis(ethylene) IrI complex [TpIr(C2H4)2] ( 1 ; Tp=hydrotris(3,5‐dimethylpyrazolyl)borate) reacts with two equivalents of aromatic or aliphatic aldehydes in the presence of one equivalent of dimethyl acetylenedicarboxylate (DMAD) with ultimate formation of hydride iridafurans of the formula [TpIr(H){C(R1)?C(R2)C(R3)O }] (R1=R2=CO2Me; R3=alkyl, aryl; 3 ). Several intermediates have been observed in the course of the reaction. It is proposed that the key step of metallacycle formation is a C? C coupling process in the undetected IrI species [TpIr{η1O‐R3C(?O)H}(DMAD)] ( A ) to give the trigonal‐bipyramidal 16 e? IrIII intermediates [TpIr{C(CO2Me)?C(CO2Me)C(R3)(H)O }] ( C ), which have been trapped by NCMe to afford the adducts 11 (R3=Ar). If a second aldehyde acts as the trapping reagent for these species, this ligand acts as a shuttle in transfering a hydrogen atom from the γ‐ to the α‐carbon atom of the iridacycle through the formation of an alkoxide group. Methyl propiolate (MP) can be used instead of DMAD to regioselectively afford the related iridafurans. These reactions have also been studied by DFT calculations.  相似文献   

3.
The known, very efficient base‐free copper(I) oxide catalyzed N‐arylation reaction performed in MeOH at room temperature for the synthesis of N‐substituted azoles and amines was extended to the heterocyclic series, i.e., we report herein the base‐free copper(I) oxide catalyzed N‐heteroarylation of 1H‐(benz)imidazole, by means of electron‐rich or electron‐deficient B‐heteroarylboronic acids or 2‐heteroaryl‐4,4,5,5‐tetramethyl‐1,3,2‐dioxaborolanes (Schemes 1 and 2). Under these conditions, N‐heteroarylated 1H‐(benz)imidazoles were obtained in good to excellent yields (Tables 1 and 2). This is the first time that 2‐heteroaryl‐4,4,5,5‐tetramethyl‐1,3,2‐dioxaborolanes were used in this type of reaction.  相似文献   

4.
Five novel pyrazole‐coupled glucosides, 1,5‐diaryl‐1H‐pyrazol‐3‐yl 2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐glucopyranosides 5a – 5e , were synthesized by the phase‐transfer catalytic reaction of 1,5‐diaryl‐1H‐pyrazol‐3‐ols 4a – 4e with acetobromo‐α‐D ‐glucose in H2O/CHCl3 under alkaline conditions, using Bu4N+Br? as catalyst. Then, glucosides 5a – 5c were deacetylated in a solution of Na2CO3/MeOH to yield the 1,5‐diaryl‐3‐(β‐D ‐glucopyranosyloxy)‐1H‐pyrazoles 6a – 6c . Their structures were characterized by 1H,1H‐COSY, 1H‐, 13C‐, and 19F‐NMR spectroscopy, as well as elemental analysis. The structures of 5d and 6c were also determined by single‐crystal X‐ray diffraction analysis. A preliminary in vitro bioassay indicated that compounds 4e and 5d exhibited excellent‐to‐medium fungicidal activity against Sclerotinia sclerotiorum at the dosage of 10 μg/ml.  相似文献   

5.
Stereoselective syntheses of (?)‐(1R,1′R,5′R,7′R)‐1‐hydroxy‐exo‐brevicomin ( 1 ) and (+)‐exo‐brevicomin ( 2 ) were accomplished from 3,4,6‐tri‐O‐acetyl‐D ‐glucal ( 5 ; Schemes 2 and 3). Chemoselective reduction, Grignard reaction, Barton? McCombie deoxygenation, and ketalization were used as key steps.  相似文献   

6.
A versatile and simple method is reported for the synthesis of bicyclic epoxide and aziridine? fused heterocycles (up to 98% yield, up to 96 : 4 er or up to 15 : 1 dr), using a tandem Michael addition/Johnson? Corey? Chaykovsky annulation approach. A new chiral (2‐bromoethyl)sulfonium reagent is described, based on an easily available chiral sulfide; it promotes or enhances stereoselectivity in the reaction.  相似文献   

7.
The kinetics of the reactions of 1,2‐diaza‐1,3‐dienes 1 with acceptor‐substituted carbanions 2 have been studied at 20 °C. The reactions follow a second‐order rate law, and can be described by the linear free energy relationship log k(20 °C)=s(N+E) [Eq. (1)]. With Equation (1) and the known nucleophile‐specific parameters N and s for the carbanions, the electrophilicity parameters E of the 1,2‐diaza‐1,3‐dienes 1 were determined. With E parameters in the range of ?13.3 to ?15.4, the electrophilic reactivities of 1 a–d are comparable to those of benzylidenemalononitriles, 2‐benzylideneindan‐1,3‐diones, and benzylidenebarbituric acids. The experimental second‐order rate constants for the reactions of 1 a – d with amines 3 and triarylphosphines 4 agreed with those calculated from E, N, and s, indicating the applicability of the linear free energy relationship [Eq. (1)] for predicting potential nucleophilic reaction partners of 1,2‐diaza‐1,3‐dienes 1 . Enamines 5 react up to 102 to 103 times faster with compounds 1 than predicted by Equation (1), indicating a change of mechanism, which becomes obvious in the reactions of 1 with enol ethers.  相似文献   

8.
The Morita? Baylis? Hillman (MBH) reactions of (4S,5R,7R,8R)‐ and (4R,5R,7R,8R)‐4‐hydroxy‐7,8‐dimethoxy‐7,8‐dimethyl‐6,9‐dioxaspiro[4.5]dec‐2‐en‐1‐ones ( 2 and 3 , resp.) with aldehydes using various catalysts were studied. A combination of Bu3P/phenol in THF was found being optimum conditions giving the corresponding MBH adducts with high diastereoisomeric ratios. After separation, each stereomerically pure isomer of the MBH adducts was subjected to hydrolysis employing 1% aq. CF3COOH (TFA) in a water bath of an ultrasonic cleaner to afford the corresponding polyhydroxylated cyclopentenones in good yields.  相似文献   

9.
Synthesis of nickel(II) complexes of meso‐aryl‐substituted azacorroles was performed by Buchwald–Hartwig amination of a dipyrrin NiII complex with benzylamine through C? N and C? C coupling. The highly planar structure of NiII azacorroles was elucidated by X‐ray diffraction analysis. 1H NMR analysis and nucleus independent chemical shift (NICS) calculation on NiII azacorrole revealed its distinct aromaticity with [17]triaza‐annulene 18π conjugation. In addition, acylation of azacorrole selectively afforded N‐ and C‐acylated azacorroles depending on the reaction conditions, showing the dual reactivity of azacorroles.  相似文献   

10.
A highly concise and stereoselective total synthesis of (5R,7S)‐kurzilactone ( 1 ) was performed by a convergent approach by means of a Jacobsen's hydrolytic kinetic resolution, a Horner? Wadsworth? Emmons reaction for the construction of the α,β‐unsaturated δ‐lactone ring system, and a highly diastereoselective Mukaiyama aldol reaction for the introduction of the formal anti‐1,3‐diol unit (Schemes 2 and 3).  相似文献   

11.
The Pd0‐mediated rapid trapping of methyl iodide with an excess amount of a heteroaryl‐substituted tributylstannane has been investigated with the aim of incorporating a short‐lived 11C‐labelled methyl group into the heteroaromatic carbon frameworks of important organic compounds, such as drugs with various heteroaromatic structures, in order to execute a positron emission tomography (PET) study of vital systems. The reaction was first performed by using our previously developed CH3I/stannane/[Pd2(dba)3]/P(o‐CH3C6H4)3/CuCl/K2CO3 (1:40:0.5:2:2:2) system in DMF at 60 °C for 5 min (conditions A), however, the reaction gave low yields for various heteroaromatic compounds. Increasing the amount of phosphine ligand (conditions B) led to a significant improvement in the yield, but the conditions were still not suitable for a range of basic heteroaromatic structures. Use of the CuBr/CsF system (conditions C) also provided a result similar to that obtained under conditions B with an increased amount of the phosphine. Thus, pyridine and related heteroaromatic compounds remained less reactive substrates. The problem was overcome by replacing the DMF solvent with N‐methyl‐2‐pyrolidinone (NMP). The reaction in NMP at 60–100 °C for 5 min using a CH3I/stannane/[Pd2(dba)3]/P(o‐CH3C6H4)3/CuBr/CsF (1:40:0.5:16:2:5) combination (conditions D) gave the methylated products in yields of more than 80 % (based on the reaction of CH3I) for all of the heteroaromatic compounds listed in this study. Thus, the combined use of NMP and an increased amount of phosphine is important for promoting the reaction efficiently. The use of this general approach to rapid methylation has been well demonstrated by the synthesis of the PET tracers 2‐ and 3‐[11C]methylpyridines by using [Pd2(dba)3]/P(o‐CH3C6H4)3/CuBr/CsF (1:16:2:5) in NMP at 60 °C for 5 min, which gives the desired products in HPLC analytical yields of 88 and 91 %, respectively.  相似文献   

12.
Novel acyclic Pd(II)‐N‐heterocyclic carbene (NHC) metallacrown ethers 5a , 5b have been synthesized. Reaction of the imidazolium salts bearing a long polyether chain with Ag2O afforded Ag‐NHC complexes, which then reacted as carbene transfer agent with PdCl2(MeCN)2 to give the desired acyclic Pd(II)‐NHC metallacrown ether complexes 5a and 5b . The 1H NMR and 13C NMR spectra show 5a and 5b exist as mixtures of cis and trans isomers in solution. The trans isomer of 5a was characterized by X‐ray diffraction, which clearly demonstrated two pseudo‐crown ether cavities in trans‐ 5a . Pd(II)‐NHC complexes 5a and 5b have been shown to be highly effective in the Suzuki‐Miyaura reactions of a variety of aryl bromides in neat water without the need of inert gas protection.  相似文献   

13.
Synthesis and Crystal Structure of a μ-Methylene-μ-hydrido-dialanate [R2Al(μ-CH2)(μ-H)AlR2]? (R = CH(SiMe3)2) tert-Butyl lithium reacts with the recently synthesized methylene bridged dialuminium compound [(Me3Si)2CH]2Al? CH2? Al[CH(SiMe3)2]2 2 in the presence of TMEDA under β-elimination; the thereby formed hydride anion is bound in a chelating manner by both unsaturated aluminium atoms forming a 3c–2e–Al? H? Al bond. The crystal structure of the product shows two independent molecules differing only slightly in bond lengths and angles, but significantly in conformation. While one of the Al2CH heterocycles deviates little from planarity with a rough C2 symmetry for the whole anion, the other one is folded with an angle of 21.1° and the arrangement of the substituents is best described by Cs symmetry.  相似文献   

14.
An efficient stereoselective total synthesis of (+)‐(4S,5S)‐muricatacin was accomplished in good yields from inexpensive, commercially available chemicals ((+)‐diethyl tartrate (DET) and undecan‐1‐ol) by utilizing Mitsunobu and Julia? Kocienski reactions, Wittig homologation, Swern oxidation, and lactonization.  相似文献   

15.
The activation of carbon–carbon σ bonds is a complementary method to access uncommon and difficult‐to‐prepare organometallic species. Herein, we describe the activation of tert‐cyclobutanols through an enantioselective insertion of a chiral rhodium(I) complex into the C? C σ bond of the cyclobutane, forming a quaternary stereogenic center and an alkyl‐rhodium functionality that initiates ring‐closure reactions. This technology provides access to a variety of substituted cyclohexane derivatives with quaternary stereogenic centers. The formation of different product families can be controlled by the employed set of reaction conditions and additives. In general, high yields and excellent enantioselectivities of up to 99 % ee are obtained.  相似文献   

16.
A general regioselective rhodium‐catalyzed head‐to‐tail dimerization of terminal alkynes is presented. The presence of a pyridine ligand (py) in a Rh–N‐heterocyclic‐carbene (NHC) catalytic system not only dramatically switches the chemoselectivity from alkyne cyclotrimerization to dimerization but also enhances the catalytic activity. Several intermediates have been detected in the catalytic process, including the π‐alkyne‐coordinated RhI species [RhCl(NHC)(η2‐HC?CCH2Ph)(py)] ( 3 ) and [RhCl(NHC){η2‐C(tBu)?C(E)CH?CHtBu}(py)] ( 4 ) and the RhIII–hydride–alkynyl species [RhClH{? C?CSi(Me)3}(IPr)(py)2] ( 5 ). Computational DFT studies reveal an operational mechanism consisting of sequential alkyne C? H oxidative addition, alkyne insertion, and reductive elimination. A 2,1‐hydrometalation of the alkyne is the more favorable pathway in accordance with a head‐to‐tail selectivity.  相似文献   

17.
A convenient and waste‐free synthesis of indene‐based tertiary carbinamines by rhodium‐catalyzed imine/alkyne [3+2] annulation is described. Under the optimized conditions of 0.5–2.5 mol % [{(cod)Rh(OH)}2] (cod=1,5‐cyclooctadiene) catalyst, 1,3‐bis(diphenylphosphanyl)propane (DPPP) ligand, in toluene at 120 °C, N‐unsubstituted aromatic ketimines and internal alkynes were coupled in a 1:1 ratio to form tertiary 1H‐inden‐1‐amines in good yields and with high selectivities over isoquinoline products. A plausible catalytic cycle involves sequential imine‐directed aromatic C? H bond activation, alkyne insertion, and a rare example of intramolecular ketimine insertion into a RhI–alkenyl linkage.  相似文献   

18.
Peripheral Bonding of Mercury(II) Iodide to Trinuclear Molybdenum-Sulfur-Dithiophosphinato Clusters: [Mo3S4(R2PS2)4HgI2] (R = Et, Pr) Reaction of Mo3S4(R2PS2)4 1 (a : R = Et, b : R = Pr) with HgI2 in THF yields the diamagnetic title complexes [Mo3S4(R2PS2)4HgI2] 3 . The crystal structure of [ 3a (H2O)] · 2 CH2Cl2 shows the complexes to consist of a triangular array of Mo atoms which are bridged by μ2? S atoms and capped by a μ3? S atom. Each of the Mo atoms is chelated by a dithiophosphinato ligand Et2PS2? and in addition two Mo atoms are bridged by a Et2PS2? ligand while the H2O molecule is bonded weakly to the third Mo atom. Thus, all Mo atoms reveal a distorted octahedral coordination sphere. HgI2 is ?peripherally”? bonded to the cluster via two S atoms, one of which belongs to a chelating ligand and the other one to the bridging ligand. Space group P1 , lattice constants a = 12.157(2), b = 15.284(3), c = 16.049(3) Å, α = 115.56(1), β = 107.35(1), and γ = 94.62(1)°; Z = 2, dcalc = 2.23 mg/mm3; 4 236 observed reflections, R = 0.068. In organic solvents complexes 3 are strong electrolytes. VT-31P NMR data suggest a stepwise dissociation of 3 with formation of [Mo3S4(R2PS2)3] +[(R2PS2)HgI2]? and elimination of the bridging ligand from the cluster.  相似文献   

19.
Chlorodibenzyltin (IV) complex with dithiomorpholinocarbamate ligand was synthesized by the reaction of dibenzyltin dichloride with dithiomorpholinocarbamate in 1:1 stoichiometry. The complex was characterized by elementary analysis, UV, BR and 1H NMR spectra. The crystal structure was determined by X‐ray single crystal diffraction study. The crystallographic data are as follows: triclinic, space group P1 , a = 0.8723 (2) ran, b = 1.099 (2) nm, c = 1.1036 (3) nm, α = 86.498 (4)°, β = 89.697 (5)°, γ = 82.807 (5)°, Z = 2, V = 1.0479 (4) nm3, Dc= 1.580 g/cm?3, μ = 1.553 mm?1, F (000) = 500, R1 = 0.0442, wR2 = 0.0974. The crystal consists of discrete molecules containing five‐coordinate tin atoms in a distorted tigonal bipyramidal configuration. The molecules are packed in the unit cell in one‐dimensional chain structure through a weak interaction between the chlorine atom and sulfur atom, the sulfur atom and one of the sulfurs of an adjacent molecule.  相似文献   

20.
Electrospray ionization of methanolic solutions of nickel(II) nitrate, 1,1′‐binaphthalene‐2,2′‐diol (BINOL), and secondary alcohols (ROH) inter alia affords monocationic complexes of the type [(BINOLate)Ni(ROH)]+, where BINOLate stands for singly deprotonated BINOL. Upon collision‐induced dissociation (CID), the mass‐selected ions undergo competing fragmentations involving loss of the alcohol ligand and expulsion of the corresponding carbonyl compound. The latter reaction leads to the hydride complex [(BINOL)Ni(H)]+ and can thus be regarded as the reversal of the reduction of ketones with metal hydrides. The possibility of the occurrence of enantioselective gas‐phase reactions is probed for combinations of chiral BINOLate ligands with chiral alkan‐2‐ols. Whereas aliphatic alkan‐2‐ols do not show pronounced chiral effects, enantioselective bond activation is observed for 1‐phenylethanol, indicating an interaction of the aromatic ring of the alkanol with the positively charged metal center.  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