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 46 毫秒
1.
This paper describes a new ring-opening-closing alternating copolymerization (ROCAC) of 2-methyl-2-oxazoline (five-membered cyclic imino ether, 1 ) with N-methyldiacrylamide ( 2 ). The reaction of a 1 : 1 monomer feed ratio proceeded without any added catalyst to give an alternating copolymer 3 having two structural units formed by ring-opening and ring-closing (cyclization). The structure of copolymer 3 was determined by 1H, 13C NMR, and IR spectroscopies. The extent of cyclization was at most 65%. The copolymerization was reasonably explained by a mechanism of propagation via zwitterion intermediates.  相似文献   

2.
A novel germanium-containing copolymer 3 a having a germanium(IV) unit and a trimethylene sulfide unit alternatingly in the main chain was prepared by the combined use of a germylene, 1,3-bis(trimethylsilyl)-1,3-diaza-2-germa(II)-indan ( 1 a ) and thietane ( 2 ). The resulting copolymer has a very high molecular weight (M?w > 106). The usage of a germylene, bis[bis(trimethylsilyl)aminato]germanium(II) ( 1 b ), as a comonomer also gave an alternating copolymer 3 b with high molecular weight in good yields. On the other hand, the reaction of a germylene, bis(N-trimethylsilyl-tert-butylaminato)germanium(II) ( c ), with thietane afforded a 1 : 1 adduct 4c having a 1-thia-2-germacyclopentane structure without producing an alternating copolymer.  相似文献   

3.
The stereochemistry of alternating copolymerization of carbon dioxide with S-(+)-cyclohexylepoxyethane ( 1 ) using a diethylzinc/water system as catalyst was investigated. The optically active copolymer, poly[(oxycarbonyloxy)-1-cyclohexylethylene], was hydrolyzed and the epoxide was transformed into the form 1,2-bis(trimethylsilyloxy)-1-cyclohexylethane ( 2 ). From the determination of optical activity of this ether, it is concluded that the ring opening of 1 in the copolymerization takes place predominantly at the methylene-oxygen linkage.  相似文献   

4.
Spontaneous copolymerization of methoxallene with 7,7,8,8‐tetracyanhoquinodimethane (TCNQ) was carried out. The obtained copolymer mainly consisted of 2,3‐polymerized methoxyallene units and TCNQ. The copolymer composition was 1 : 1 irrespective of the monomer feed ratios, suggesting that an alternating copolymerization occurred. UV‐Vis and ESR spectroscopic studies supported that the active species of the copolymerization was a zwitterion generated through a single electron transfer reaction from methoxyallene to TCNQ.  相似文献   

5.
The copolymerization of vinyl chloride (VC) and methyl acrylate (MA) in the presence of ethylaluminium compounds (C2H5AlCl2, (C2H5)2AlCl, and (C2H5)3Al) at low ethylaluminium compound (EAC)/MA mole ratios was investigated. An alternating copolymer was produced in this reaction when an excess of VC in the initial monomer feed was used. The addition of dibenzoyl peroxide (BPO) to the systems containing EAC resulted in an increase of the alternating copolymer yield. In polymerization systems containing EAC resulted in an increase of the alternating copolymer yield. In polymerization systems containing EAC combined with VOCl3 an enhancement of the alternating copolymer yield and formation of VC-rich copolymers were observed. In the polymerization system with (C2H5)3Al? VOCl3 a VC-rich copolymer was the main product. It was concluded that VC-rich copolymers are formed in the random radical copolymerization which occurs when most of EAC is complexed by the alternating copolymer chain. The structure of alternating and VC-rich copolymers was studied in detail by means of 13C NMR spectroscopy.  相似文献   

6.
Copolymerization of cyclohexene oxide (CHO) and sulfur dioxide (SO2) was conducted at 80°C by using a variety of catalysts both in the presence and absence of solvent. Aromatic tertiary amines such as pyridine gave an alternating copolymer of CHO and SO2 in the presence of solvents such as 1,2-dimethoxyethane. The alternating copolymer obtained was a white thermoplastic which could readily be thermoformed. It was soluble in acetone and chloroform but insoluble in water and hexane. From the reaction of the equimolar mixture of CHO, SO2 and pyridine at 20°C, an 1 : 1 : 1 adduct of CHO, SO2 and pyridine (see formula 2 ) could be isolated. This adduct was confirmed to be the active species for the present copolymerization.  相似文献   

7.
Carbon dioxide was copolymerized with propylene oxide (PO, 2-methyloxirane) in the presence of a catalyst system based on alkylmetal compounds (MtRn, Mt = Zn, Cd, Al) and pyrogallol (PG) at mole ratios 1:1, 2:1, and 3:1, respectively. The reaction was found to produce an alternating copolymer poly(propylene carbonate), [poly(oxycarbonyloxypropylene)], using Zn and Cd containing catalysts and non-alternating copolymers with Al containing catalysts. For the system ZnR2/PG (mole ratio 2 : 1), the most active system, the influence of substituents R at the zinc atom, as well as effects of solvents and complexing agents on the copolymer yield and intrinsic viscosity were investigated. The results are discussed in terms of an anionic coordinative copolymerization mechanism.  相似文献   

8.
The copolymerization of methyl acrylate (MA) and isobutylene (IB) in the presence of Lewis acids (EtAlCl2, Et2AlCl, Et3Al, AlCl3, and ZnCl2) at low Lewis acid/MA mole ratio was investigated. EtAlCl2 and Et2AlCl were found to initiate the spontaneous reaction. An alternating copolymer was produced in this reaction when an excess of IB in the initial monomer feed was used. The copolymerization in the presence of Et3Al, AlCl3, and ZnCl2 did not proceed spontaneously and was initiated by dibenzoyl peroxide (BPO). In this case MA-rich copolymers are formed even in systems containing a large excess of IB in the monomer feed. The addition of BPO to systems containing ethylaluminium chlorides strongly diminishes the tendency towards alternating propagation. It was concluded that the mode of initiation has a significant influence on the copolymer composition. The alternating copolymerization by EtAlCl2 was studied in detail in order to determine the influence of the catalyst concentration, monomer feed ratio, reaction temperature and time on the monomer conversion, copolymer composition, molecular weight and tacticity.  相似文献   

9.
The 1H NMR spectra of random and alternating copolymers of styrene and methyl methacrylate, ethyl methacrylate, butyl methacrylate, and octyl methacrylate were analyzed. In the first approximation, the one-parameter statistics suggested for the interpretation of these spectra is in agreement with the experimental data. The way in which the copolymer is synthesized and the copolymerization mechanism following therefrom markedly affect the magnitude of the parameter of coisotacticity σ: for all comonomer pairs under study, σ is higher if an organometallic catalyst is used (alternating copolymerization) than in the case of a radical initiator (statistical copolymerization). With increasing number of carbon atoms in the n-alkyl alcohol residue, σ decreases to a limiting value.  相似文献   

10.
1,3-Bis(4-aminophenyl)imidazolidinetrione dihydrochloride ( 1 ) was synthesized and used as starting material for the synthesis of strictly alternating copolymers. Fully aromatic copolymers containing parabanic acid and amide moieties ( 7a, b ) were found to exhibit high glass temperatures and good thermal stability. Meltable copolymers with parabanic acid and amide moieties ( 7d–j ) could be obtained by introducing flexible aliphatic spacers. The use of an unusually short, but symmetrically branched spacer was particularly effective. The strictly alternating, thermally stable copolymer 10 was synthesized from 1 and pyromellitic dianhydride. By using the trimethylsilyl ester of polyamic acid 11 , the imidization temperature could be lowered by 100°C.  相似文献   

11.
Ring-opening alternating copolymerization of maleic anhydride and epichlorohydrin was carried out at 70°C with rare earth coordination catalysts composed of Nd[(RO)2 PO2]3 (R = CH3(CH2)3CH(C2H5)CH2? )and trialkylaluminium. Anlaysis of end-groups showed that the copolymer chain contains one ? OH and one ? CH?CH? CO? CH2CH(CH3)2 end-group. IR, UV-Vis, 1H NMR and GPC results imply that a catalyst-maleic anhyride complex is formed in the initiation step that the ring-opening copolymerization proceeds via coordinate insertion mechanism accompanied with chain-transfer.  相似文献   

12.
The formation of a charge-transfer complex between 1-vinylindole ( 1a ) and maleic anhydride (MA) was detected by 1H NMR analysis, and the corresponding Keq was measured (Keq=0,251 mol in CDCl3 at 25°C). It could be shown that this complex can be involved both in the initiation and propagation steps of the 1a -MA copolymerization by carrying out spontaneous copolymerization experiments and studying the influence of the feed composition on the conversion and [η] of the copolymer. The structure of the 1a -MA copolymer, investigated by 13C NMR spectroscopy, was found to be complicated due to the formation of indoline rings along the chains by an attack of the propagating MA radical at position 2 in the indole ring. As a consequence an excess of MA, with respect to a 1:1 mole ratio of monomer feed, was usually found in the copolymer. The mechanism of the cyclization reaction was clarified by copolymerizing MA with 2-methyl-1-vinylindole ( 1b ), 3-methyl-1-vinylindole ( 1c ), or 2,3-dimethyl-1-vinylindole ( 1d ). An alternating copolymer with MA (mole ratio 1:1 of monomeric units), without irregularities, was obtained only by employing 1d as a comonomer.  相似文献   

13.
The stereochemistry of alternating copolymerization of carbon dioxide with S-(–)-3-phenyl-1,2-epoxypropane ( 5 ) using a diethylzinc/water system as catalyst was investigated. The optically active copolymer, poly[(oxycarbonyloxy)-1-benzylethylene] ( 6 ), was hydrolyzed and the epoxide unit was transformed into the form of 1,2-bis(trimethylsilyloxy)-3-phenylpropane ( 8 ). From the determination of optical activity of this ether, it is concluded that the ring opening of 5 in the copolymerization takes place predominantly at the methylene-oxygen linkage. The substituent effect on the ring opening mode of oxiranes is also discussed.  相似文献   

14.
The copolymerization of 1-(2-hydroxyethyl)aziridine ( 1 ) as nucleophilic monomer with β-butyrolactone ( 2 ) as electrophilic monomer without initiator in solution at 45°C was investigated. Copolymers were characterized by IR and 1H NMR spectroscopy. By elemental analyses it could be shown that their compositions depend on the monomer ratio in the feed. From the copolymer composition and 1H NMR spectra it was possible to suggest four probable copolymer structures.  相似文献   

15.
The zwitterionic copolymerization of 2-methyl-2-oxazoline (MOX) as nucleophilic monomer with β-butyrolactone (BUL) as electrophilic monomer was investigated in bulk and in solution (CH3CN) at 45°C. The copolymer composition was around 1,5/1,0 (BUL/MOX) as was established by 1H NMR. 1H and 13C NMR spectroscopy were used to identify the copolymers. The IR spectroscopy supported the NMR results. On the other hand, the copolymers behave as polyelectrolytes, according to viscosity determinations. A copolymerization mechanism through a zwitterion species is suggested.  相似文献   

16.
Carbon dioxide was copolymerized with methyloxirane and 7-oxabicyclo[4.1.0]heptane using triethylaluminium as a catalyst component. Factors to control the copolymer composition were sought on the basis of an assumed copolymerization mechanism. Compared to nonpolar solvents, ethereal solvents were found to increase the carbon dioxide content in the copolymer obtained by the triethylaluminium/water system, giving a fraction with almost alternating sequence. The addition of a Lewis base, especially treiphenylphosphine, to triethylaluminium was also useful to get copolymers of high carbon dioxide content. The coincidence of these results with the expectation from the assumed copolymerization scheme clearly indicates that the copolymerization proceeds through a coordinated anionic mechanism.  相似文献   

17.
To clear up the detailed mechanism of the alternating copolymerization of styrene (St) and maleic anhydride (MAn) concerning the initiation species, the propagation species, the tendency of the chain transfer reaction as well as the directional tendency of the reaction between copolymer radicals and monomer complexes, the role of the charge transfer complexes, characterization of end groups and additional donor effects were examined. The equilibrium constant of the St/MAn (1 : 1) complex was determined to be 0.31 by NMR spectroscopy, that suggested considerable amounts of complexes existing in the system. As expected, a small quantity of initiator (14C-azobisisobutyronitrile (AIBN)) was incorporated into the St/MAn copolymer. Chlorine atoms were scarcely incorporated into the copolymer synthesized in CCl4 with AIBN or benzoyl peroxide (BPO) as an initiator. Hence the copolymerization was considered to be induced only by attacke of initiator radicals to the monomer or the complexes, contrary to the usual conception of telomerization. When the electron donor monomer was added to the system, the terpolymerization could be treated as a copolymerization of the two complexes, i.e., St/MAn and Donor/MAn. By adding naphthalene the rate maximum point shifted from higher concentration of MAn to the equivalent concentration of St and MAn. Degradative chain transfer to N.N-dimethylaniline was observed, confirming the existance of poly-MAn radicals. It was suggested from these results that the charge transfer complex and uncomplexed MAn took part in the copolymerization of St and MAn. This was proved kinetically. The whole mechanism was discussed.  相似文献   

18.
The asymmetric induction radical copolymerization of styrene (St) with maleic anhydride (MAn) was studied in various solvents with different dielectric constants in the presence of lecithin as a chiral surfactant. The observed specific rotation of the resulting alternating copolymer decreased with an increase in the dielectric constant of the solvent used. It was concluded that the asymmetric induction depends mainly upon the incorporation of MAn within the reversed lecithin micelles.  相似文献   

19.
The present paper describes the ternary catalyst system consisting of Group II metal halides, AlEt3, and CCl4, which initiatesatroom temperature the polymerizations of ethylene (60 atmospheres), styrene, vinyl acetate, methyl methacrylate, acrylonitrile, and isobutyl vinyl ether. As the component of Group II metal halide, ZnCl2, ZnI2, CdCl2, MgCl2, and BeCl2 were effective. This catalyst system induced the cationic polymerization of vinyl ether. On the other hand, the copolymer composition curve of the styrene-methacrylate copolymerization with this catalyst system was close to that of free radical copolymerization. In addition, this catalyst system has also been characterized by the 1:1 alternating copolymerization of ethylene with vinyl acetate at room temperature.  相似文献   

20.
Real‐time infrared spectroscopy was used to monitor the high‐speed copolymerization of N‐substituted alkyl maleimide (MI) and vinyl ether (VE) mixtures upon UV exposure, with or without photoinitiator. The monomer feed composition was shown to play a decisive role on the polymerization kinetics of each monomer. When VE is in excess, the two monomers disappear at similar rates to yield an alternating copolymer. When MI is in excess, the copolymerization and the MI homopolymerization occur simultaneously to yield a copolymer with isolated VE units. The reaction scheme proposed to account for the kinetic data is based on the homopolymerization of a donor‐acceptor complex, with the VE∗︁ radicals acting as the main propagating species. The kinetic chain length was evaluated from quantum yield measurements performed under controlled initiation conditions and found to be on the order of 103. The propagation rate constant of the VE/MI copolymerization was calculated to be twice that of the MI homopolymerization.  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