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 31 毫秒
1.
In this study, polymers of the MMA/Styrene/BMA three‐component system were synthesized through either soapless semibatch emulsion copolymerization or soapless batch emulsion copolymerization technique. The optimal monomer feed flow rate was determined from the interphase partition laws, monomer reactivity ratios, and three or four times of iterative experimental procedures through semibatch emulsion copolymerization. As a result, the instantaneous composition of polymers could also be effectively controlled to get the desired final products. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3253–3269, 2000  相似文献   

2.
Formaldehyde oxime and three O‐alkyl derivatives were examined as potential imine monomers. Formaldehyde oxime spontaneously polymerized below 60 °C and did not polymerize above 60 °C (ceiling temperature), even in the presence of free‐radical or cationic initiators. The O‐benzoyl derivative was isolated as the cyclic trimer but could not be converted into the monomeric form. Formaldehyde O‐benzyloxime was synthesized and isolated. Attempted homopolymerizations in the presence of free‐radical initiators only led to oligomers, whereas with cationic initiators only cyclic trimer was obtained. Copolymerizations with appropriate vinyl monomers and free‐radical and anionic initiators yielded only low molecular weight polymers. Cationic copolymerizations gave higher molecular weights and polymer yields, but the polymers containing appreciable amounts of imine function had very low molecular weights. We conclude that the polymerizability of imines is extremely sensitive to the substitution pattern: imines with only a substituent on nitrogen are unstable and readily polymerize, whereas imines with more substituents generally do not polymerize. Electron‐withdrawing substituents are more favorable to polymerizability. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1866–1872, 2000  相似文献   

3.
Radical homopolymerizations and copolymerizations of 3‐[tris(trimethylsilyloxy)silyl]propyl methacrylate (SiMA) in supercritical CO2 were investigated. The homopolymer was obtained in CO2 with a good yield. It was essentially insoluble in pure CO2 at less than 500 bar at 65 °C but was soluble in a mixture of CO2 and its monomer (10 w/v %) at 352 bar. The copolymerizations of SiMA with methyl methacrylate, 1,1‐dihydroperfluorooctyl methacrylate, and styrene with various monomer feed ratios were also examined in supercritical CO2 and in bulk, and the reactivity ratios were determined. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3100–3105, 2000  相似文献   

4.
Novel hyperbranched polymers were synthesized in a high yield without gelation through the free‐radical alternating copolymerization of an AB/B′ (allyloxy maleic acid/maleic anhydride) system, in which group B and monomer B′ both could only alternately polymerize with group A. The arm number of the produced highly branched polymers was equal to the product of the linear chain length and the probability of pendent B groups being growing centers. The molecular weight of these novel hyperbranched polymers increased with increasing initiator concentration and prolonged polymerization times. The AB/B′ system, used as described, provides a new general methodology for highly branched and functional polymers. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3074–3085, 2000  相似文献   

5.
2-Hydroxyethyl methacrylate (HEMA) and styrene (S) have been copolymerized in a 3 mol · L−1N,N′-dimethylformamide (DMF) solution using 2,2′azobis (isobutyronitrile) (AIBN) as an initiator over a wide composition and conversion range. From low-conversion experiments and 1H-NMR analysis, the monomer reactivity ratios were determined according to the Mayo–Lewis terminal model. The comparison of the obtained results with those previously reported for copolymerization in bulk and in toluene reveals a relatively small but noticeable solvent effect that can be qualitatively explained by the bootstrap model. Cumulative copolymer composition as a function of conversion is satisfactorily described by the integrated Mayo–Lewis equation; overall copolymerization rate increases with increasing the HEMA/S ratio, and individual monomer conversion is closely related to the monomer molar fraction in the feed. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2941–2948, 1999  相似文献   

6.
Butyl acrylate (BA) and methyl methacrylate (MMA) have been copolymerized in a 3 mol/L benzene solution using 2,2′-azobis(isobutyronitrile) (AIBN) as initiator over a wide composition and conversion range. The overall copolymerization parameter kp/kt1/2 and the composition of the copolymer formed have been measured as a function of conversion. Theoretical values of the coupled parameter kp/kt1/2 calculated from the implicit penultimate unit model and those of cumulative copolymer composition, determined from the Mayo—Lewis terminal model, have been correlated with those experimentally obtained. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35 : 1961–1965, 1997  相似文献   

7.
A variety of new polymeric materials ranging from soft rubbers to hard, tough, and brittle plastics were prepared from the cationic copolymerization of regular soybean oil, low saturation soybean oil (LoSatSoy oil), or conjugated LoSatSoy oil with styrene and divinylbenzene initiated by boron trifluoride diethyl etherate (BF3 · OEt2) or related modified initiators. The relationship between the dynamic mechanical properties of the various polymers obtained and the stoichiometry, the types of soybean oils and crosslinking agents, and the different modified initiators was investigated. The room‐temperature storage moduli ranged from 6 × 106 to 2 × 109 Pa, whereas the single glass‐transition temperatures (Tg) varied from approximately 0 to 105 °C. These properties were comparable to those of commercially available rubbery materials and conventional plastics. The crosslinking densities of the new polymers were largely dependent on the concentration of the crosslinking agent and the type of soybean oil employed and varied from 74 to 4 × 104 mol/m3. The Tg increased and the intensity of the loss factor decreased irregularly with an increase in the logarithmic crosslinking densities of the polymers. Empirical equations were established to describe the effect of crosslinking on the loss factor in these new polymeric materials. The polymers based on conjugated LoSatSoy oil, styrene, and divinylbenzene possessed the highest room‐temperature moduli and Tg 's. These new soybean oil polymers appear promising as replacements for petroleum‐based polymeric materials. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 2721–2738, 2000  相似文献   

8.
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   

9.
The synthesis of a diblock copolymer of styrene and vinyl acetate (VAC), PS‐b‐PVAC, was performed by successive photoinduced charge‐transfer polymerization (CTP) under UV irradiation. A novel amphiphilic diblock copolymer of PS‐b‐PVA then was obtained by the hydrolysis of the diblock copolymer PS‐b‐PVAC with sodium ethoxide as a catalyst. Both of them were characterized by Fourier transform infrared, H NMR, and gel permeation chromatography in detail. The effect of the solvents on the CTP and the kinetics of the CTP are discussed. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 914–920, 2000  相似文献   

10.
Core–shell lattices with a polystyrene core and a polystyrene/butyl acrylate shell with more than 40% solid contents were produced using a combination of sodium dodecyl maleate hemiester as anionic surfactant and styrenic block copolymer of butylene oxide and ethylene oxide as nonionic surfactant. Stable lattices able to resist rather high concentrations of electrolytes can be obtained, provided a careful protocol of the addition of surfactants is used. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2251–2262, 1999  相似文献   

11.
The phase‐transfer catalyzed graft copolymerization of methyl methacrylate onto polychloroprene was carried out using tricaprylylmethylammonium chloride as a phase‐transfer catalyst in a two‐phase system of an aqueous Na2S2O8 solution and toluene at 55 °C under a nitrogen atmosphere. The initial rate of graft copolymerization was expressed as the combined terms of quaternary onium cation and peroxydisulfate anion in the aqueous phase rather than the fed concentrations of catalyst and Na2S2O8. The observed initial rate of graft copolymerization was used to analyze the graft copolymerization mechanism with a cycle phase‐transfer initiation step in the heterogeneous liquid–liquid system. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3543–3549, 2000  相似文献   

12.
We investigate the effect of the presence of reactive impurities during copolymerization on mathematical model discrimination and determine if impurities adversely affect the model discrimination process, depending on the criteria and response(s) used. Nonlinear mechanistic process models are employed in a multiresponse statistical discrimination scenario. The investigation also considers, via several case studies, the extremely interesting question of whether impurity effects could lead to incorrect choices about the copolymerization mechanism itself (terminal versus penultimate model). The answers to the above questions are that reactive impurities do affect the model discrimination process and, what is more, they may mislead researchers as to the choice of the copolymerization mechanism. The latter unexpected result is further illustrated with experimental data from the literature. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2319–2332, 2000  相似文献   

13.
Enantioselective, alternating copolymerizations of carbon monoxide with styrene, dicyclopentadiene, and methylcyclopentadiene dimer were carried out with a palladium catalyst modified by 1,4‐3,6‐dianhydro‐2,5‐dideoxy‐2,5‐bis(diphenyl phosphino)‐L ‐iditol. Chiral diphosphine was proven to be effective at enantioselective copolymerization. In the copolymers, some of the second double bonds of alternating poly(1,4‐ketone) were carbonylated. Optical rotation, elemental analysis, and spectra of 1H NMR, 13C NMR, and IR showed that the copolymers had isotactic, alternating poly(1,4‐ketone) structures. An oxidant and an organic acid were the promoters of the copolymerization. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2919–2924, 2000  相似文献   

14.
An investigation was carried out to examine the conduction mechanism in poly(pyromellitic‐4,4‐diphenyl sulphone) [PPDS] films prepared by cast method. The electrical properties were measured in aluminum/polymer/aluminum structure over the temperature range (30–100 °C). At low field region, Ohm's law was obeyed whilst at fields high enough to cause injection of charge carriers, the electrodes supplied all the carriers and Schottky effect mechanism was observed as a prevail one. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 2507–2514, 2000  相似文献   

15.
A series of hydrophobic poly(oxypropylene) (POP)‐backboned and hydrophilic poly(oxyethylene)‐backboned amidoacids and imidoacids were prepared through the reaction of poly(oxyalkylene) diamines and trimellitic anhydride (TMA) under mild conditions. The synthesized copolymers were characterized with nuclear magnetic resonance and Fourier transform infrared. Their ability to lower the water surface tension and toluene/water interfacial tension was measured and correlated with the hydrophobic/hydrophilic balance with multiple sodium carboxylate functionalities. The specific POP2000/TMA copolymers, consisting of a 2000 g/mol POP segment and multiple amidoacid functionalities, enabled the demonstration of a strong surfactant tendency and a critical micelle concentration at 0.1 wt %. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 646–652, 2006  相似文献   

16.
Disclinations in the fingerprint‐like cholesteric texture of a natural polymer derivative, cyanoethyl chitosan, were observed and studied by both polarizing optical microscopy and scanning electron microscopy. In the latter technique, permanganic etching was developed for this cholesteric texture to increase the contrast. A special mechanism for cholesteric phase etching is suggested and discussed. Some χ, λ, and τ disclinations and domain walls were observed. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 980–986, 2000  相似文献   

17.
A series of novel polyesters containing dioxane moieties in their main chains were synthesized by the bulk polycondensation of trans‐2,5‐bis‐(hydroxy‐ methyl)‐1,4‐dioxane with various aliphatic dicarboxylic acid chlorides. The obtained polyesters, analyzed by differential thermal analysis, possessed crystallinity, the melting point of which exhibited a weak odd–even effect on the methylene unit number and a small decreasing trend with an increase in the methylene unit number. These properties were compared with those of similar polyesters bearing cyclohexane moieties, and it was found that the rigidity of the dioxane moiety plays an important role in enhancing the effective packing of the corresponding polymers. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2536–2542, 2000  相似文献   

18.
Exact equations and several computer programs were developed for use in studies on copolymerizations carried to high conversion when the concentration of one of the monomers (A) remains constant. Simple ACSL® and DESIRE® programs are described for simulating such copolymerizations, and their output was used to test programs and procedures that were developed to evaluate monomer reactivity ratios for such copolymerization systems. Based on an integrated form of the copolymer equation, Excel® and Fortran programs were developed for evaluating monomer reactivity ratios from information about initial monomer compositions, copolymer compositions, and the fractions of the second monomer (B) that reacted. A graphical procedure for evaluating monomer reactivity ratios from such data was also developed. A previous program developed for calculating information about monomer sequence distributions in copolymers was modified so that it would apply to copolymerization at high conversion when the concentration of one monomer remains constant. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1118–1128, 2000  相似文献   

19.
Products of the radical dispersion copolymerization of methacryloyl‐terminated poly(ethylene oxide) (PEO) macromonomer and styrene were separated and characterized by size exclusion chromatography (SEC), full adsorption‐desorption (FAD)/SEC coupling and eluent gradient liquid adsorption chromatography (LAC). In dimethylformamide, which is a good solvent for PEO side chains but a poor solvent for polystyrene (PS), amphiphilic PS‐graft‐PEO copolymers formed aggregates, which were very stable at room temperature even upon substantial dilution. The aggregates disappeared at high temperature or in tetrahydrofuran (THF), which is a good solvent for both homopolymers and for PS‐graft‐PEO. FAD/SEC procedure allowed separation of homo‐PS from graft‐copolymer and determination of both its amount and molar mass. Effective molar mass of graft‐copolymer was estimated directly from the SEC calibration curve determined with PS standards. Presence of larger amount of the homo‐PS in the final graft‐copolymer products was also confirmed with LAC measurements. The results indicate that there are at least two or maybe three polymerization loci; namely the continuous phase, the particle surface layer and the particle core. The graft copolymers are produced mainly in the continuous phase while PS or copolymer rich in styrene units is formed mostly in the core of monomer‐swollen particles. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2284–2291, 2000  相似文献   

20.
Poly(sulfone‐arylate) was synthesized in a reaction between dihydroxy polysulfone prepolymers and either diphenyl terephthalate or terephthaloyl chloride. The dihydroxy polysulfone prepolymers had molecular weights of 2000 and 4000 g/mol. The polymerization with diphenyl terephthalate was carried out at high temperature (280 °C) in the presence of a catalyst, whereas the polymerization with terephthalic chloride was conducted in solution at low temperature in the presence of an acid acceptor. High‐molecular weight copolymers (ηinh ~ 0.60 dL/g) could be obtained through both methods. The copolymers were characterized by Fourier transform infrared spectroscopy, nuclear magnetic resonance, DMA, and differential scanning calorimetry measurements and were found to exhibit high Tg values. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 3904–3913, 2009  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