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.
The kinetic behavior of the 60Co-initiated copolymerization at 25°C of styrene with vinyl acetate at 1100 and 2000 rad/hr was studied. As in the case of thermal and photochemical copolymerizations of these monomers, the growing chains are particularly rich in styrene units, and the overall rate is affected by a diluent effect due to the vinyl acetate monomer. However, in the case of the radiation copolymerization, this effect is partially counterbalanced by an increase of the initiation rate with the vinyl acetate concentration; the polymerization rate curve shows a maximum at a vinyl acetate molar fraction of 0.25. This effect is due to the very different free radical yields of these two monomers. The experimental results may be understood on the basis of a kinetic scheme which involves an energy transfer process from the excited vinyl acetate molecules to the styrene monomer and a termination reaction of the growing chains by very short styrene radicals when the mixture is rich in vinyl acetate.  相似文献   

2.
Emulsion polymerization of vinyl benzoate and its copolymerization with vinyl acetate or styrene are described. The effect of the potassium persulfate initiator, and the sodium lauryl sulfate emulsifier concentration on the rate of vinyl benzote homopolymerization and the molecular weight of the polymers was determined. In copolymerization with vinyl benzoate, both comonomers, vinyl acetate and styrene, decrease the initial polymerization rate. With increasing amounts of styrene in the comonomer mixture the polymerization rate increases but with vinyl acetate an opposite effect is observed. Reactivity ratios of copolymerizations were determined. For the vinyl benzoate [M1]-styrene [M2] comonomer system a r1 = 0.03 and a r2 = 29.58 and for vinyl benzoate [M1]-vinyl acetate [M2], a r1 = 1.93 and a r2 = 0.20 was obtained. From the vinyl benzoate-styrene reactivity ratios the Qe parameters were calculated.  相似文献   

3.
Perfluorocyclopropene undergoes free-radical copolymerization with ethylene, isobutylene, cis- and trans-2-butene, vinyl acetate, methyl vinyl ether, vinyl chloride, styrene, acrylonitrile, tetrafluoroethylene, vinyl fluoride, and vinylidene fluoride. The copolymerization proceeds most readily with electron-rich olefins such as methyl vinyl ether (to yield a 1:1 copolymer), but conditions were found to give copolymers with electron-deficient olefins such as tetrafluoroethylene and vinylidene fluoride. Copolymers with methyl vinyl ether, tetrafluoroethylene, vinyl fluoride, and vinylidene fluoride were examined in detail. Evidence is presented that the perfluorocycloproply ring is incorporated intact into the copolymer and can be subsequently isomerized to a perfluoropropenyl unit by heating at 200–300°C.  相似文献   

4.
This paper discusses monomer reactivity ratios in various radiation- and redox-initiated graft copolymerizations. The polymers studied were polyethylene, cellulose acetate, poly(vinyl chloride), polytetrafluoroethylene, poly(vinyl alcohol), and poly(methyl methacrylate); the comonomer mixtures were styrene–acrylonitrile, methyl acrylate–styrene, acrylonitrile–methyl acrylate, and vinyl acetate–acrylonitrile. The polymer–comonomer mixture systems were so chosen as to permit study of both homogeneous and heterogeneous systems. The homogeneous systems included systems of low and high viscosity. The heterogeneous systems included both polymers swollen by the comonomer mixture and polymers not swollen by the comonomer mixture. None of the homogeneous grafting systems studied showed deviations from the normal copolymerization behavior under a variety of experimental conditions. Monomer reactivity ratios in graft copolymerization were the same as the values in nongraft copolymerization. The heterogeneous systems in which the polymer was swollen by the comonomer mixture yielded grafted copolymer compositions which were the same as those in nongraft copolymerization. The heterogeneous grafting system polytetrafluoroethylene/styrene–acrylonitrile showed deviations from normal copolymerization behavior at low degrees of grafting when the reaction was only on the polymer surface. The behavior became normal at higher degrees of grafting when the system approaches that in which the polymer is swollen by the comonomers. In all reaction systems, it was found that the use of radiation to initiate the reaction does not in any way affect the copolymerization behavior of the two monomers in a comonomer pair.  相似文献   

5.
The vinyl ester of dehydrogenated abietic acid has been homopolymerized, copolymerized with vinyl chloride, vinyl acetate, and butadiene and terpolymerized with styrene and acrylonitrile. Both in the homopolymerization and the copolymerization, the vinyl ester of dehydroabietic acid has given lower molecular weight polymers than were obtained with the vinyl ester of tetrahydroabietic acid. Polymers containing the vinyl ester of dehydroabietic acid can be readily crosslinked with peroxide.  相似文献   

6.
Poly[(vinyl alcohol)-co-(vinyl amine)] was prepared by copolymerization of vinyl acetate and N-vinyl forma-mide, followed by a two-step alkaline hydrolysis in which the vinyl acetate groups were hydrolyzed to vinyl alcohol groups, and the N-vinyl formamide groups, to amino groups. This copolymer was used as the stabilizer for the emulsion polymerization of styrene using persul-fate-ion initiator. The latexes were cationic; the particles were relatively uniform and of 100–500 nm diameter. The presence of amino groups was shown by electro-phoresis and Kjeldahl analysis. Comparison with sodium lauryl sulfate showed that the mechanism of particle nucleation was apparently entry of radicals into fine emulsion droplets rather than the conventional entry of radicals into emulsifier micelles.  相似文献   

7.
A nonaqueous potentiometric direct titration method was used to determine the composition diagrams for the copolymerization of maleic anhydride with styrene and with vinyl acetate in methyl ethyl ketone at 50°C. The data were analyzed using nonlinear least-squares minimization routines to fit composition equations for the terminal, penultimate, and complex models directly. The applicability of each model to both systems were evaluated statistically. The penultimate model was found to best describe both systems, although in the case of the maleic anhydride/vinyl acetate system this was only a small improvement over the terminal model. Although significant comonomer complexation occurs in both systems, the complex model did not provide statistically significant improvement in fit to the data compared with the terminal model.  相似文献   

8.
Polymerizations of methyl methacrylate (MMA), vinyl acetate, acrylonitrile or styrene (St) initiated by methyl derivatives of Ti(IV) are consistent with a radical process. This conclusion is supported by tacticities of the resulting polymers which are close to those for conventional radical samples and show that propagation obeys a Bernoullian statistics. The major evidence is provided by statistical MMA-St copolymerization which clearly exhibits radical reactivity ratios.  相似文献   

9.
1-Methylcyclopropene (MCP) copolymerizes rapidly with acrylic and vinyl monomers to form soluble, high molecular weight products containing enchained cyclopropane rings. The high electron availability in the cyclopropene double bond promotes one-to-one alternating copolymerization with sulfur dioxide, maleic anhydride, acrylic acid, acrylonitrile, dialkyl fumarates and acrylic esters. Nonalternating copolymers are obtained with vinyl chloride and vinyl acetate, and attempted copolymerization fails entirely with styrene, α-methylstyrene and isoprene. This pattern of copolymerization reactivity resembles that of highly compressed ethylene. Methylcyclopropene copolymers have high glass temperatures in spite of the small size of the MCP unit. The combination of high Tg and small size allows preparation of copolymers with high Tg having a wide range of ductilities and cohesive energy densities.  相似文献   

10.
Copolymers of methyl α-cyanocrotonate with styrene, acrylonitrile, and vinyl acetate were prepared in bulk by free radical initiation. The copolymerization parameters were determined for each pair by several methods. The basic properties, that is, intrinsic viscosity, solubility, melting range, and glass transition temperature of the obtained copolymers, were determined.  相似文献   

11.
The aqueous solution copolymerization of maleic anhydride(MAn), vinyl acetate (VAc) and styrene(St) was carried out by using redox system as initiator. Aqueous MAn solution PH should be adjusted and controlled between 6 and 7 by slow addition of proper amount of aqueous NaOH solution. Initiator used for copolymerization was redox system of K2S2Os-NaHSO3. The polymerization temperature was maintained at 70-72℃ for approximately 7 hours. After polymerization reaction, water in mixture of polymerization was evaporated by means of water bath. The mixture was dried in vacuum desiccator and purified. The product of copoly(MAn-VAc-St) is yellow powder.  相似文献   

12.
Photografting copolymerization of maleic anhydride (MAH) and styrene (St) onto LDPE film was investigated by using a one-step method, and further thermally induced grafting copolymerization of them was carried out by using a two-step method. Regarding the photografting copolymerization of MAH/St binary monomer system, both conversion percentage (CP) and grafting efficiency (GE) increased with raising the content of MAH in the monomer feed. In addition, the content of MAH in the grafted copolymers also increased with increasing the fraction of MAH in the monomer feed. The formation of LDPE-g-P(MAH-co-St) grafted film was identified by FTIR and ESCA spectroscopy. In the case of grafting copolymerization of MAH/St by the two-step method, grafting copolymerization proceeded slowly compared with the non-grafting copolymerization. The apparent activation energy (Ea) for the non-grafting copolymerization in the solution and the grafting copolymerization on LDPE film was 24 and 82 kJ/mol, respectively, which were noticeably lower than those of MAH/vinyl acetate (MAH/VAC) binary monomer system under the similar grafting conditions. These data of Ea explained why the grafting copolymerization of styrene/MAH took place faster than that of MAH/VAC binary monomer system. The composition of the grafted copolymer chains was largely affected by the composition of the monomer feeds; however, the composition of the non-grafted copolymers nearly remained at 1/1 even in systems with largely different MAH/styrene ratios in monomer feeds. It is indicated that the non-grafting copolymerization proceeded predominantly following alternating copolymerization, but the grafting copolymerization performed random copolymerization.  相似文献   

13.
The copolymerization of vinyl acetate (VAc) and vinyl 2-ethylhexanoate (V2EH) via miniemulsion polymerization is being investigated with the goal of better copolymer composition control. The reaction kinetics and final particle sizes were compared for homopolymerizations and copolymerizations of the two monomers. Enhanced polymerization rates were seen for the miniemulsion polymerizations over their conventional counterparts despite a decreased number of particles. This was attributed to the influence of the reactive surfactant sodium dodecyl allyl sulfosuccinate and low water solubility of the V2EH monomer. Increased nucleation of miniemulsion droplets was sought by incorporation of small amounts of polymer into the droplets. This was first shown to be effective for styrene miniemulsion polymerizations and was subsequently successfully extended to the VAc/V2EH system  相似文献   

14.
The crosslinking performance of the unsaturated hyperbranched polyester poly(allyloxy maleic acid‐co‐maleic anhydride) (MAHP) was investigated with copolymerizations of three different monomers: styrene, vinyl acetate, and methyl methacrylate. Both styrene and vinyl acetate afforded interpenetrating‐polymer‐network copolymer gels. The gels exhibited crosslink density gradients through the polymer matrices on a macroscopic level, and density maximums were concentrated around the MAHP moieties. The heterogeneity of the gels is briefly discussed in terms of a modified two‐phase model, where one phase consists of an elastic part of low crosslinking density and the other phase consists of an inelastic dendritic part with a highly condensed bond density. Unlike the two‐phase model developed by Choquet and Rietsch, the modified two‐phase model takes into account that both phases swell in good solvents. Unlike copolymerizations employing styrene or vinyl acetate, the copolymerization of MAHP with methyl methacrylate afforded noncrosslinked starbranched copolymers that consisted of a MAHP core from which long poly(methyl methacrylate) branches were protruding. The different behaviors of the copolymerizations of the three monomers used in this study can rationally be explained by their different reactivity ratios with maleic end groups of MAHP. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 964–972, 2001  相似文献   

15.
<正> 用大分子单体与小分子单体共聚是七十年代初才出现的合成接枝共聚物的一种新方法。通过共聚合反应而不是接枝反应同时形成主干及支链。这种接枝共聚物由于形成支链的大分子单体是预先合成的,其分子量分布较窄,又可调节控制,所以合成的接枝共聚物支链长短比较均一,副反应较少,链结构比较明确,因此也易于表征。  相似文献   

16.
Abstract

The change in relative reactivity in the cationic copolymerization of 2-chloroethyl vinyl ether and styrene derivatives was investigated with various catalysts and solvents. p-Methoxystyrene, p-methylstyrene, and a-methyl-styrene were used as styrene derivatives. The styrene content in the co-polymer increased when a polar solvent and/or a strong catalyst was used. The change of relative reactivity in the copolymerization of 2-chloroethyl vinyl ether with styrene derivatives was much greater than that in the copolymerization between vinyl ethers or styrene derivatives. When nitro-ethane was used as a solvent, not only the polarity but also the nucleophilicity influenced the copolymer composition. The results were discussed by two energies, Eπ and Ers, which are measures of complex formation between monomer and carbonium ion, and stabilization energy in the transition state, respectively.  相似文献   

17.
The kinetics of homopolymerization of N-vinyl-3(5)-methylpyrazole, its copolymerization with vinyl acetate, 2-hydroxyethyl methacrylate, and acrylic acid, as influenced by the reaction medium, were studied. The properties of the resulting (co)polymers were determined.  相似文献   

18.
Mechanical and structural-rheological properties of oxidized bitumen modified with styrene butadiene-styrene copolymers side vinyl groups in their butadiene fragment were studied. A comparative analysis of the modified and unmodified bitumen composites was performed.  相似文献   

19.
Vinyl acetate forms a complex with manganese tris(acetylacetonate). A scientific hypothesis, according to which the use of manganese tris(acetylacetonate) as an initiator would lead to the increase in the relative activity of vinyl acetate in copolymerization with more active monomers, is suggested. To verify the hypothesis experimentally, an integrated study of the kinetics of binary copolymerization of vinyl acetate with N-vinyl (N-vinylsuccinimide and N-vinyl-3(5)methylpyrazole) and acrylic (2-hydroxyethyl methacrylate) monomers in various media is carried out. It is demonstrated on the basis of an analysis of copolymerization constants of the monomers under study that the relative activities of monomers in fact become closer. This leads to a change of the microstructure of polymers and results in macrochains with improved alteration of monomer units.  相似文献   

20.
Atom transfer radical homo- and copolymerization of styrene and methyl acrylate initiated with CCl3-terminated poly(vinyl acetate) macroinitiator were performed at 90°C in the presence of nanoclay (Cloisite 30B). Controlled molecular weight characteristics of the reaction products were confirmed by GPC. It was shown that nanoclay slightly decreased the rate of styrene polymerization, while it significantly enhanced the rate of methyl acrylate polymerization and its copolymerization with styrene. The reactivity ratios of the monomers in the presence and in the absence of nanoclay were calculated (r St = 1.002 ± 0.044, r MA = 0.161 ± 0.018 by extended Kelen-Tudos method and r St = 1.001 ± 0.038, r MA = 0.163 ± 0.016 by Mao-Huglin method), confirming that the presence of nanoclay has no influence on monomer reactivity. The enhancement in the homopolymerization rate of methyl acrylate as well as its copolymerization rate with styrene was attributed to nanoclay effect on the dynamic equilibrium between active (macro)radicals and dormant species. Dipole moments of the monomers were successfully used to predict structure of the polymer/clay nanocomposites prepared via in situ polymerization.  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