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.
This research deals with comparison of the activity of various Rh catalysts in the polymerization of monosubstituted acetylenes and the effect of various amines used in conjunction with [Rh(nbd)Cl]2 in the polymerization of phenylacetylene. A zwitterionic Rh complex, Rh+(nbd)[(η6‐C6H5)B?(C6H5)3] ( 3 ), was able to polymerize phenylacetylene ( 5a ), t‐butylacetylene ( 5b ), N‐propargylhexanamide ( 5c ) and n‐hexyl propiolate ( 5d ), and displayed higher activity than the other catalysts examined, that is [Rh(nbd)Cl]2 ( 1 ), [Rh(cod)(Oo‐cresol)]2 ( 2 ), and Rh‐vinyl complex ( 4 ). Monomers 5a and 5c polymerized virtually quantitatively or in fair yields with all these catalysts, while monomer 5b was polymerizable only with catalyts 3 and 4 . Monomer 5d did not polymerize in high yields with these Rh complexes. The catalytic activity tended to decrease in the order of 3 > 4 > 2 > 1 . Although polymerization of 5a did not proceed at all in toluene with [Rh(nbd)Cl]2 alone, it smoothly polymerized in the presence of various amines as cocatalysts. The polymerization rate as well as the molecular weight of polymer depended on the basicity and steric bulkiness of amines. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 4530–4536, 2005  相似文献   

2.
N-Carbazolylacetylene (CzA) was polymerized in the presence of various transition metal catalysts including WCl6, MoCl5, [Rh(NBD)Cl]2, and Fe(acac)3 to give polymers in good yields. The polymers produced with W catalysts were dark purple solids and soluble in organic solvents such as toluene, chloroform, etc. The highest weight-average molecular weight of poly(CzA) reached about 4 × 104. In the UV–visible spectrum in CHCl3, poly(CzA) exhibited an absorption maximum around 550 nm (εmax = 4.0 × 103 M−1 cm−1) and the cutoff wavelength was 740 nm, showing a large red shift compared with that of poly(phenylacetylene) [poly(PA)]. Poly(CzA) began to lose weight in TGA under air at 310°C, being thermally more stable than poly(PA) and poly[3-(N-carbazolyl)-1-propyne]. Poly(CzA) showed a third-order susceptibility of 18 × 10−12 esu, which was 2 orders larger than that of poly(PA). © 1998 John Wiley & Sons, Inc. J. Polym. Sci. A Polym. Chem. 36: 2489–2492, 1998  相似文献   

3.
Some polyacetylene derivatives containing an amine functional group were prepared by the polymerization of propargylamine (PA) and 1,1-diethylpropargylamine (DEPA) with various transition metal catalysts. In the polymerization of PA, Mo-based catalysts were more effective than that of W-based catalysts, and organoaluminum compounds, especially EtAlCl2, were found to be very effective cocatalysts. In the polymerization of DEPA, Mo-and W-based catalyst systems showed a similar catalytic activity. The polymerization easily proceeded in polar solvents such as nitrobenzene and DMF as well as nonpolar aromatic solvents such as chlorobenzene, toluene, etc. The resulting poly(PA) and poly(DEPA) were insoluble in organic solvents regardless of polymerization catalysts but the polymers were found to be stable to air oxidation. Thermogravimetric analyses and thermal transitions of poly(PA) and poly(DEPA) were also studied. © 1992 John Wiley & Sons, Inc.  相似文献   

4.
Iron(II) dichloride complexes bearing a tridentate nitrogen donor ligand were investigated for the homopolymerization of tert‐butyl acrylate after activation with methylaluminoxane. Two new complexes were synthesized, 2,6‐bis[1‐(cyclohexylimido)ethyl]pyridine iron chloride (FeCl2) and 2,6‐bis[1‐(isopropylimido)ethyl]pyridine FeCl2, and the single‐crystal X‐ray structure of the latter one was determined. Turnover frequencies of the catalysts during polymerization ranged from 36 to 241 cycles/mol/h. The obtained polymers exhibited weight‐average molecular weights ranging from 24,000 to 438,500 g/mol, and the molar mass distribution varied between 1.8 and 4.3. Activity of the catalytic system and the molar mass of the polymer are influenced by the ligand structure of the complexes as well as other polymerization conditions such as monomer concentration and temperature. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 1380–1389, 2003  相似文献   

5.
A wide range of cyclopentadienylnickel compounds catalyse the reaction of phenylacetylene under solvent‐free conditions, giving a mixture of cyclotrimers, linear oligomers and poly(phenylacetylene). No reaction is observed in the case of internal acetylenes. Cyclotrimer formation is favoured by the presence of cyclopentadienylnickel catalysts bearing a chloro substituent at nickel. A reduction in reaction temperature results in lower conversion but favours linear oligomer and polymer formation. The main effect of the presence of solvent, regardless of whether it is potentially coordinating (toluene) or not (n‐octane), is to suppress almost completely reactions catalysed by nickelocene.  相似文献   

6.
Silica‐supported titanium(IV) chloride is readily reduced by Mashima and co‐workers' reagent (1‐methyl‐3,6‐bis(trimethylsilyl)‐1,4‐cyclohexadiene) to afford materials active in ethylene polymerisation without need of aluminum alkyl cocatalyst.  相似文献   

7.
8.
The immobilization of [Rh(cod)OCH3]2 (cod = cycloocta‐1,5‐diene) on mesoporous molecular sieves MCM‐41 provides the first inorganic‐type hybrid catalyst, which affords heterogeneous polymerization of phenylacetylene and its ring‐substituted derivatives, – 2‐fluorophenylacetylene, 4‐fluorophenylacetylene, and 4‐pentylphenylacetylene – into readily isolable high‐molecular‐weight (w from 50 000 to 180 000) substituted polyvinylenes of high cis‐transoid structure. The activity of this catalyst is compared with that of homogeneous catalyst [Rh(cod)OCH3]2.  相似文献   

9.
1-Ethynylcyclohexene, an acetylene derivative having cyclohexenyl substituent, was polymerized by various W- and Mo-based catalysts. WCl6-EtAlCl2 catalyst system was found to be very effective for this polymerization. The effects of the monomer-to-catalyst mol ratio, the initial monomer concentration, the temperature, and the cocatalysts for the polymerization of 1-ethynylcyclohexene by WCl6 were investigated. The catalytic activity of Mo-based catalysts was found to be similar to that of W-based catalysts. The polymer structure was identified to have a conjugated polymer backbone carrying a cyclohexenyl substituent. The resulting polymers were light-brown powder and completely soluble in aromatic and halogenated hydrocarbon solvents such as chlorobenzene, benzene, chloroform, carbon tetrachloride, etc. Studies of the thermal properties and morphology of poly(1-ethynylcyclohexene) were also carried out. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
End‐functionalized poly(phenylacetylene)s were synthesized by the polymerization of phenylacetylene (PA) using the well‐defined palladium catalysts represented as [(dppf)PdBr(R)] {dppf = 1,1′‐bis(diphenylphosphino)ferrocene}. The Pd catalysts having a series of R groups such as o‐tolyl, mesityl, C(Ph)?CPh2, C6H4o‐CH2OH, C6H4p‐CN, and C6H4p‐NO2 in conjunction with silver triflate polymerized PA to give end‐functionalized poly(PA)s bearing the corresponding R groups in high yields. The results of IR and NMR spectroscopies and MALDI‐TOF mass analyses proved the introduction of these R groups at one end of each polymer chain. The poly(PA) bearing a hydroxy end group was applied as a macroinitiator to the synthesis of a block copolymer composed of poly(PA) and poly(β‐propiolactone) moieties. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

11.
12.
The polymerization of 1-methoxy-1-ethynylcyclohexane (MEC) was carried out by various transition metal catalysts. The catalysts MoCl5, MoCl4, and WCl6 gave a relatively low yield of polymer (≤ 16%). The catalytic activity of Mo-based chloride catalyst was greater than that of W-based chloride catalyst. However, catalyst tungsten carbene complex (I) gave a larger molar mass and higher yield in the presence of a Lewis acid such as AlCl3 than in the absence of a Lewis acid. The activity of the tungsten carbene complex was obviously affected by Lewis acidity. The catalyst PdCl2 was a very effective catalyst for the present polymerization and gave polymers in a high yield. The structure of the resulting poly(MEC) was identified by various instrumental methods as a conjugated polyene structure having an α-methoxycyclohexyl substituent. The poly(MEC)s were mostly light-brown powders and completely soluble in various organic solvents such as tetrahydrofuran (THF), chloroform (CHCl3), ethylacetate, n-butylacetate, dimethylformamide, benzene, xylene, dimethylacetamide, 1,4-dioxane, pyridine, and 1-methyl-2-pyrrolidinone. Thermogravimetric analysis showed that the polymer started to lose mass at 125°C and that maximum decomposition occurred at 418°C. The x-ray diffraction diagram shows that poly(MEC) has an amorphous structure. © 1997 John Wiley & Sons, Inc.  相似文献   

13.
Iron catalysts were evolved for more active transition metal‐catalyzed living radical polymerization through design of the ligands. In situ introduction of P–N chelate‐ligands, consisting of hetero‐coordinating atoms [phosphine (P) and nitrogene (N)], onto FeBr2 effectively catalyzed living radical polymerization of methyl methacrylate (MMA) in conjunction with a bromide initiator, where the monomer‐conversion reached over 90% without dropping the rates and the molecular weights of obtained PMMAs were well controlled. The benign effects of the “hetero‐chelation” were demonstrated by comparative experiments with homo‐chelate ligands (P–P, N–N), model compounds of the composed coordination site, and the combinations. We successfully achieved an isolation of iron complex with a P–N ligand [FeBr2(DMDPE); DMDPE: (R)‐N,N‐dimethyl‐1‐(2‐(diphenylphosphino)phenyl)‐ethanamine], which was superior to the conventional catalyst [FeBr2(Pn‐Bu)2] with respect to controllability and activity, especially at the latter stage. The catalyst was almost quantitatively removed by water washing after polymerization. It was also effective for living polymerization of styrene. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 6819–6827, 2008  相似文献   

14.
Summary: The polymerizations of ethyl vinyl ether, n‐butyl vinyl ether and isobutyl vinyl ether were investigated with a series of pyridine bis(imine) complexes of iron(II ) and cobalt(II ) in the presence of methylaluminoxane. The cobalt catalysts showed much higher activity and produced higher molecular weight polymers than their iron analogues. Both catalyst systems produced predominantly atactic polymers. There were no specific trends in the activity and the polymer molecular weight, according to the steric bulk around the metal center.

The iron(II ) and cobalt(II ) catalysts used here.  相似文献   


15.
The reaction of 2‐[bis(2‐methoxy‐phenyl)phosphanyl]‐4‐methyl‐benzenesulfonic acid (a) and 2‐[bis(2′,6′‐dimethoxybiphenyl‐2‐yl)phosphanyl]benzenesulfonic acid (b) with dimethyl(N,N,N,N′‐tetramethylethylenediamine)‐palladium(II) (PdMe2(TMEDA)) leads to the formation of TMEDA bridged palladium based polymerization catalysts ( 1a and 1b ). Upon reaction with pyridine, two mononuclear catalysts are formed ( 2a and 2b ). These catalysts are able to homopolymerize ethylene and also copolymerize ethylene with acrylates or with norbornenes. With ligand b , high molecular weight polymers are formed in high yields, but higher comonomer incorporations are obtained with ligand a .

  相似文献   


16.
Molybdenum chloride (MoCl5 or 1a ) and tungsten chloride (WCl6 or 1b )/phenyllithium (PhLi)/triisobutylaluminum (iBu3Al) systems were found to be quite effective for controlling the anionic polymerization of methyl methacrylate (MMA), affording high molecular weight poly(methyl methacrylate)s (PMMAs; number‐average molecular weight > 100,000) with narrow molecular weight distributions (weight‐average molecular weight/number‐average molecular weight < 1.25) quantitatively at 0 °C for 1 h in toluene. Matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS) analyses of PMMAs obtained with the 1a and 1b /organolithium (RM; n‐butyllithium, PhLi)/iBu3Al systems revealed that the initiation of MMA with the systems occurred by a nucleophilic attack of H? to the monomer. In addition, the MALDI‐TOF MS analyses indicated that the presence of iBu3Al was responsible for the controlled polymerization by improving the uniformity of the polymerization with respect to initiation and termination and by preventing a backbiting reaction. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4302–4315, 2002  相似文献   

17.
Anionic polymerization of methyl methacrylate (MMA) initiated with late transition‐metal halides [manganese chloride (MnCl2), iron dichloride (FeCl2), iron trichloride (FeCl3), cobalt chloride (CoCl2), or nickel bromide (NiBr2)]/organolithium [nButyllithium (nBuLi) or phenyllithium (PhLi)]/triisobutylaluminum (iBu3Al) systems is described. Except for the system with NiBr2, the polymerizations of MMA afforded narrow molecular weight distribution poly(methyl methacrylate)s (PMMAs) with high molecular weights in quantitative yields at 0 °C in toluene. Matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS) analyses of the PMMAs obtained by the systems with FeCl2, FeCl3, and CoCl2 revealed that the polymers had hydrogen (H) at both chain ends. Accordingly, the reaction of the transition‐metal halides with the organolithium in the presence of iBu3Al should result in the formation of transition‐metal hydride [M‐H]? species, which was nucleophilic enough to initiate the MMA polymerization. Because the presence of a six‐membered cyclic structure resulting from backbiting was confirmed from the MALDI‐TOF MS analyses of the PMMA obtained with the metal halide (FeCl2, FeCl3, or CoCl2)/organolithium systems in the absence of iBu3Al, the introduction of H at the ω‐chain end indicated that iBu3Al should prevent the backbiting. However, the MnCl2/nBuLi/iBu3Al initiating system gave PMMAs bearing H at the α chain end and six‐membered cyclic structure at the ω chain end. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 1962–1977, 2003  相似文献   

18.
Poly(phenylacetylene)s containing pendant phosphorescent iridium complexes have been synthesized and their electrochemical, photo‐ and electroluminescent properties studied. The polymers have been synthesized by rhodium‐catalyzed copolymerization of 9‐(4‐ethynylphenyl)carbazole (CzPA) and phenylacetylenes (C∧N)2Ir(κ2O,O′‐MeC(O)CHC(O)C6H4C?CH‐4) (C∧N = κ2N,C1‐2‐(pyridin‐2‐yl)phenyl (IrppyPA) or κ2N,C1‐2‐(isoquinolin‐1‐yl)phenyl (IrpiqPA)). In addition, organic poly(phenylacetylene)s with pendant carbazole groups have been synthesized by rhodium‐catalyzed copolymerization of CzPA and 1‐ethynyl‐4‐pentylbenzene. Complex (C∧N)2Ir(κ2O,O′‐MeC(O)CHC(O)Ph) (IrpiqPh; C∧N = 2‐(isoquinolin‐1‐yl)phenyl‐κ2N,C1) was prepared and characterized. While the copolymers of the Irppy series were weakly phosphorescent, those of the Irpiq series displayed at room temperature intense emissions from the carbazole (fluorescence) and iridium (phosphorescence) emitters, being the latter dominant when the spectra were recorded using polymer films. Triple layer OLED devices employing copolymers of the Irpiq series or the model complex IrpiqPh yielded electroluminescence with an emission spectra originating from the iridium complex and maximum external quantum efficiencies of 0.46% and 2.99%, respectively. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3744–3757, 2010  相似文献   

19.
Polymerizations of vinyl ethers are carried out with (α‐diimine)nickel(II ) catalysts in the presence of methylaluminoxane. Effects of structural variations of the ligand on the activities of catalysts and polymer microstructure are described. The catalysts prepared by changing the bulkiness of ligand substituents in the ortho aryl position result in no specific trends terms of the yield and molecular weight of polymer. Poly(vinyl ether)s are atactic regardless of the structure of the catalyst used.

  相似文献   


20.
Some new Rh(I) and Ir(I) complexes of the types [(COD)M(LL)]ClO4 and [(COD)MCl]2 [COD = cyclooctadiene; M = Rh, Ir; LL = 1,1′-bis(diphenylphosphino)ferrocene (DPPF), 1-diphenylphosphino-2-(N,N-dimethylamino)methylferrocene (FcNP), 1,6-diferrocenyl-2,5-diazahexane (FcNN)] were prepared, and their catalytic activities toward polymerization of phenyl acetylene were examined. The rhodium complexes proved to be very effective catalysts to yield highly stereoregular polyphenylacetylene (cis-transoidal-PPA) in high yields under mild conditions. The number-average molecular weight (M n) of the PPA obtained is in the range of 19,000–33,000 and the weight-average molecular weight (M ω) is in the range of 47,000–95,000. Comparative studies revealed that of various catalysts employed, the cationic mononuclear [Rh(FcNN)(COD)]ClO4 complex exhibited the best results to give exclusively the cis-transoidal-PPA (cis content ∼100%) with the highest molecular weight (M n = 33,340) in the highest chemical yield (94%). Other reaction parameters such as the softness of the ligand, the solvent, the relative amount of catalyst, and the reaction temperature were also investigated to find that all these factors played crucial roles. The iridium systems worked better for the trimerization rather than polymerization to yield 1,3,5-triphenybenzene as major product. © 1996 John Wiley & Sons, Inc.  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