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.
Methods of preparing new monomers, 2-vinyl and 2-isopropenyloxazoles and 2-isopropenyl-1,3,4-oxadiazoles are described. New methods were developed to synthesize monomers containing an isoxazole or a thiazole ring. Radical homopolymerization and copolymerization with styrene of these monomers were carried out by using AIBN as an initiator. Monomer reactivity ratios r1, r2 and Alfrey-Price Q–e values were determined by the Fineman-Ross and the Mayo-Lewis methods. The localization energy of the β-carbon was calculated by a HITAC-5020 computer, and the monomer reactivity is discussed in terms of Lβ.  相似文献   

2.
Two single oxygen-bridged heterobimetallic oxides of Al(III) with group 4 metals (Ti, Hf) have been prepared. The reaction of LAlMeOH (1) [L = CH(N(Ar)(CMe))2, Ar = 2,6-iPr2C6H3] with dimethylmetallocenes of Ti and Hf in toluene (80 degrees C) and ether (room temperature), respectively, resulted in the formation of LAl(Me)(mu-O)M(Me)Cp2 [M = Ti (2), Hf (3)] in moderate to good yield. Compounds 2 and 3 were characterized by elemental analysis, IR, NMR (1H and 13C), EI-MS, and single-crystal X-ray structural analysis. Furthermore, compound 2 showed good catalytic activity in ethylene and styrene homopolymerization, while compound 3 is less active in ethylene polymerization. The styrene polymerization yields atactic polystyrene.  相似文献   

3.
This communication reports the styrene homopolymerization behavior and ethylene-styrene copolymerization behavior of the covalently linked bimetallic constrained geometry catalyst (mu-CH2CH2-3,3'){(eta5-indenyl)[1-Me2Si(tBuN)](TiMe2)}2 (Ti2), which is the first single-site catalyst that effects not only styrene homopolymerization with high activity, but also efficient ethylene-styrene copolymerization over a broad styrene composition range (0-76% at 20 degrees C, 1.0 atm ethylene pressure). In styrene homopolymerization, a 50x increase in polymerization activity is achieved with Ti2 vs the mononuclear analogue, Ti1, using an identical trityl borate cocatalyst and polymerization conditions. In ethylene + styrene copolymerization, Ti2 enchains approximately 20% more styrene than Ti1 under identical reaction conditions. 13C NMR spectroscopy indicates that greater than two consecutive styrene units are enchained in the copolymer backbone produced by Ti2 + Ph3C+B(C6F5)4-. End group analysis of the styrene homopolymer produced by Ti2 + Ph3C+B(C6F5)4- suggests that 1,2-regiochemistry is installed in approximately 50% of the initiation steps. This unusual microstructure is believed to be related to the bimetallic catalyst structure.  相似文献   

4.
As in the case of vinylhydroquinone (I), its alkyl-substituted derivative, 2-methyl-5-vinylhydroquinone (II) was found to copolymerize with methyl methacrylate by tri-n-butylborane in cyclohexanone at 30°C. II was prepared from the O,O′-bisether compound, 2-methyl-5-vinyl-O,O′-bis(1′-ethoxyethyl)hydroquinone (III). The monomer reactivity ratios (M2 = II) were determined to be r1 = 0.37 and r2 = 0. No homopolymerization proceeded under the same conditions. Ordinary free-radical initiators, such as azobisisobutyronitrile and benzoyl peroxide, were not effective in the homopolymerization of II. 1:1 Copolymers were obtained from II and maleic anhydride by using tri-n-butylborane as an initiator. The copolymers exhibited no definite melting range and decomposed at 370–375°C endothermally (DSC). The polymerization behavior of III was also investigated. Although tri-n-butylborane did not initiate the homopolymerization of the monomer, azobisisobutyronitrile was capable of initiating the homopolymerization and copolymerization of III. The monomer reactivity ratios (M1 = styrene) were determined to be r1 = 0.83 and r2 = 0.18. The ratios gave the following Q and e values; Q = 0.15 and e = ?2.2.  相似文献   

5.
Summary: The radical homo‐ and copolymerization of styrene ( 1 ) and diethyl fumarate (DEF, 2 ) in the presence of methylated β‐cyclodextrin (β‐CD) in water is described. It has been shown for the first time that homopolymerization of CD‐complexed DEF and its copolymerization with CD‐complexed styrene occur readily in aqueous solution. In the absence of CD, or in organic solvents, the homopolymerization of DEF is strongly retarded.

  相似文献   


6.
Ethylene/styrene copolymerization catalyzed by half-titanocenes [Cp′TiX3 (Cp′ = cyclopentadienyl; X = halogen, alkyl etc.), linked half-titanocenes, and modified half-titanocenes expressed as Cp′TiX2(Y) (Y = anionic donor ligand)] have been reviewed. Results in the syndiospecific styrene polymerization using Cp′TiX2(Y)–MAO catalysts have also been summarized. The activity and molecular weight in the resultant (co)polymer (and styrene incorporation) are highly affected by ligand (Cp′ and Y). It has been suggested that the cationic Ti(IV) plays a role in the copolymerization whereas the neutral Ti(III) plays a role in the homopolymerization.  相似文献   

7.
Density functional theory has been employed to study the homogeneous catalytic copolymerization of styrene with carbon monoxide.The copolymerization reaction is catalyzed by Pd(Ⅱ) coordinated with 2,2’-bipyridine,a conventional nitrogen-containing bidentate ligand with achiral C2vsymmetry.The chain propagation mechanism for the alternating copolymerization as well as the side reactions,including multiple insertions of CO and homopolymerization of styrene,has been investigated.This study focused exclusively on regioisomerism and stereoisomerism.We have demonstrated that the strictly alternating copolymerization is kinetically and thermodynamically favored over the side reactions(i.e.,multiple insertions of CO and homopolymerization of styrene).The regiochemistry study indicates the 2,1 type.Furthermore,the stereochemistry study shows that the syndiotactic conformation is preferred over the isotactic or atactic conformations.  相似文献   

8.
New cyclopropane-containing optically transparent polymers were prepared by radical homopolymerization of p-(2-methoxycarbonylcyclopropyl)styrene and by its binary copolymerization with styrene and methyl methacrylate. The composition and structure of these macromolecular compounds were determined, and their main service characteristics were evaluated.  相似文献   

9.
The copolymerization of ethylene with styrene by Cp*TiCl2(N=CtBu2) (Cp* = C5Me5) took place in a living manner in the presence of methylaluminoxane (MAO) cocatalyst, although the homopolymerization of neither ethylene nor styrene proceeded in a living manner. Both the cyclopentadienyl fragment (Cp') and the anionic donor ligand (X) in Cp'TiCl2(X) directly affect the copolymerization behavior, the catalytic activities, as well as the styrene incorporation; only the above set showed a living copolymerization. No styrene repeating units were observed in the resultant poly(ethylene-co-styrene)s, suggesting that a certain degree of the styrene insertion inhibited the chain transfer in this catalysis.  相似文献   

10.
Four new substituted styrene derivatives carrying lactam rings (2-pyrrolidone or 2-piperidone) in para position have been synthesized, namely 4-(2-oxo-3-methylene-pyrrolidinyl)styrene, 4-(2-oxo-3-methylene-piperidinyl)styrene, 4-(p-styryl)-2-pyrrolidone, and 4-(p-styryl)-2-piperidone. Their homopolymerization and copolymerization with styrene, methyl methacrylate, and acrylic acid have been considered. By ring opening of the side lactam groups, the homopolymers are transformed into the corresponding poly aminocarboxylic acids.  相似文献   

11.
Styrene/maleic anhydride (MA) copolymerization was carried out using benzoyl peroxide (BPO) and 2,2,6,6‐tetramethyl‐1‐piperidinyloxy (TEMPO). Styrene/MA copolymerization proceeded faster and yielded higher molecular weight products compared to styrene homopolymerization. When styrene/MA copolymerization was approximated to follow the first‐order kinetics, the apparent activation energy appeared to be lower than that corresponding to styrene homopolymerization. Molecular weight of products from isothermal copolymerization of styrene/MA increased linearly with the conversion. However products from the copolymerization at different temperatures had molecular weight deviating from the linear relationship indicating that the copolymerization did not follow the perfect living polymerization characteristics. During the copolymerization, MA was preferentially consumed by styrene/MA random copolymerization and then polymerization of practically pure styrene continued to produce copolymers with styrene‐co‐MA block and styrene‐rich block. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2239–2244, 2000  相似文献   

12.
Two new monomers with pendent 1,3,4-oxadiazoline-5-thione structures were prepared. Homopolymerization of 2-isopropenyl-1,3,4-oxadiazoline-5-thione (V) and copolymerization with styrene (M1)(r1 = 0.02, r2 = 1.56, Q = 4.12, e = 1.06) were examined. Further, 2-(4-methacryloylaminophenyl)-1,3,4-oxadiazoline-5-thione (VIII) having a phenylcarbamoyl group between the isopropenyl and 1,3,4-oxadiazoline-5-thione ring as a spacer was also synthesized and polymerized. The resultant polymers were allowed to react with N-protected α-amino acids such as Z-AlaOH, Z-LeuOH and Z-PheOH by the DCC method. The polymers containing amino acids thus obtained were reacted with ethyl glycinate to give the corresponding dipeptides in excellent yields without racemization.  相似文献   

13.
Esters or carbonates of N‐hydroxypyridine‐2‐thione (Barton esters) were appended to either carboxymethyl or hydroxypropyl cellulose. Irradiation of the cellulose bound Barton esters in monomer initiated free radical graft copolymerization with minimal concomitant homopolymerization. Grafting of styrene to carboxymethyl cellulose was accompanied by backbone cleavage. The hydroxypropyl spacer group minimized backbone degradation; styrene, acylamide and N‐isopropyl acrylamide could be grafted to hydroxypropyl cellulose in tetrahydrofuran solution. Treatment of Barton carbonate modified hydroxypropyl cellulose with styrene in the presence of TEMPO afforded corresponding TEMPO adducts, which can be used to promote the controlled radical graft polymerization of styrene. Grafts were analyzed independently after hydrolysis of the cellulose backbone.  相似文献   

14.
以丁基锂为引发剂、四氢呋喃为溶剂 ,在氮气气氛和 - 78℃下进行苯乙烯 (ST)阴离子聚合 ,并加入 3 溴噻吩作为终止剂 ,从而合成出末端含噻吩基的聚苯乙烯 (PST Thp) ,它可以进行氧化偶联聚合 ,是一类新的大分子单体 .在无水三氯化铁作用下 ,PST Thp可以进行均聚反应或与噻吩进行共聚反应 ,共聚反应生成主链刚性、支链分子量分布窄的聚 (噻吩 接枝 苯乙烯 ) (PThp g ST) ,凝胶渗透色谱、红外光谱和吸收光谱证实接枝共聚物的生成  相似文献   

15.
A density functional theory (B3LYP) computational study of the ethylene–styrene copolymerization process using meso‐Et(H4Ind)2Zr(CH3)2 as the catalyst is presented. The monomer insertion barriers in meso species are evaluated and compared with previously obtained barriers in rac diastereoisomers. Differences related to ethylene homopolymerization and ethylene–styrene copolymerization activities as well as styrene incorporation into the copolymer are found between the meso and rac diastereoisomers. Nevertheless, a migratory insertion mechanism seems to hold for both diastereoisomeric species. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4752–4761, 2006  相似文献   

16.
氮气保护下二氯甲烷中铬(III)四苯基卟啉衍生物在-40℃与亚碘酰苯反应,分离得氧配位铬(V)四苯基卟啉配合物:O=Cr(V)TPP(Cl)PhI,O=Cr(V)TPP(N~30PhI,O=Cr(V)TPP(p-CH~3O-C~6H~4O)(1/2)PhI。已经元素分析、可见、红外、顺磁、核磁和质谱法结构表征。这些配合物能氧化苯乙烯,环己醇,环己烯和环己烷,可作为细胞色素P-450模拟体系的活性中间体。  相似文献   

17.
A series of monocyclopentadienyl titanium complexes containing a pendant amine donor on a Cp group ( A = CpTiCl3, B = CpNTiCl3, C = CpNTiCl2TEMPO, for Cp = C5H5, CpN = C5H4CH2CH2N(CH3)2, and TEMPO = 2,2,6,6‐tetramethylpiperidine‐N‐oxyl) are investigated for styrene homopolymerization and ethylene–styrene (ES) copolymerization. When activated by methylaluminoxane at 70 °C, complexes with the amine group ( B and C ) are active for styrene homopolymerization and afford syndiotactic polystyrene (sPS). The copolymerizations of ethylene and styrene with B and C yield high‐molecular weight ES copolymer, whereas complex A yields mixtures of sPS and polyethylene, revealing the critical role that the pendant amine has on the polymerization behavior of the complexes. Fractionation, NMR, and DSC analyses of the ES copolymers generated from B and C suggest that they contain sPS. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1579–1585, 2010  相似文献   

18.
The radiation induced graft copolymerization of m-isopropenyl-α,α-dimethyl benzyl iso-cyanate (m-TMI) and styrene onto polypropylene was carried out. The extent of grafting increased with increasing amount of styrene and with increased radiation dose. A graft load of 180% was obtained by immersing a 50 kGy pre-irradiated film in a monomer solution containing 25 mol % m-TMI and 75 mol % styrene. The graft copolymer is suitable for covalently binding nonpolymerizable stabilizers with a suitable nucleophilic moiety. In this work the isocyanate moiety of the graft copolymer was allowed to react with 4-amino-2,2,6,6-tetramethyl piperidine, a hindered amine light stabilizer. Fourier trans-formed infrared spectroscopy confirmed the formation of an urea moiety. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
Styrene-terminated poly(oxyethylene) macromonomers (SOE) with narrow molecular weight distribution and quantitative styrene monofunc-tionality were synthesized. In homopolymerization of SOE, conversion of monomer to polymer was shown to be low in spite of high consumption of the vinyl groups of the SOE molecules. Free-radical copolymer-ization of the macromonomer with methyl methacrylate and styrene occurred smoothly, as opposed to homopolymerization. Cumulative copolymer composition and total conversion were determined from the conversions of macromonomer and comonomer (by weight changes) and by proton NMR of the copolymer. The monomer reactivity ratios were found to be ra = 0.06 and rb = 2.0 for the copolymerization of SOE macromonomer (a) with methyl methacrylate (b). In this case the macromonomer exhibited considerably lower reactivity than predicted from its low molecular weight model compound. The monomer reactivity ratios estimated for SOE and styrene were ra = 0.86 and rb = 1.20. The reactivity of SOE was comparable to, but somewhat lower than, styrene. The graft copolymers were used as activators in the halogen displacement reaction, and it was found that their catalytic activity depends on copolymer composition and chemical structure.  相似文献   

20.
A mathematical model for describing the particle size distribution (PSD) in emulsion copolymerization system is developed by analogy to that in emulsion homopolymerization system as proposed by Lichti and co-workers. By use of the appropriate combinations of the kinetic parameters of the comonomers, the complicated equations for copolymerization systems can be reduced to simpler equations identical to those of homopolymerization systems. The two calculation examples, styrene–methyl methacrylate and styrene-butadiene systems, are given to demonstrate the applicability of the proposed theory. The conditions for producing bimodal PSD from a seeded emulsion polymerization are discussed.  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