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.
Dialkyl iodomethylmalonates in the presence of tetrabutylammonium iodide initiate the polymerization of methyl methacrylate, but not of methyl acrylate or acrylonitrile. Typically, at 60°C in 1,3-dimethyltetrahydro-2-1H-pyrimidone (DMPU) as the solvent, poly(methyl methacrylate (PMMA)) is obtained in the number-average molecular weight range of 2 000 to 8 000, the molecular weight distribution being fairly narrow (ratio of weight- to number-average molecular weights w/n 1.2–1.3).  相似文献   

2.
Organomanganate reagents [R3Mn]Li+ (R = Bu, Me) were found to polymerize methyl methacrylate in the presence of potassium tert‐butylate. A conversion of the tacticity of the resulting poly(methyl methacrylate)s from heterotactic (mr = 54%) to isotactic (mm = 58%) was observed upon changing the R group of the initiator from Bu to Me. The addition of triisobutylaluminium was found to efficiently control w and w/n of the resulting polymers.  相似文献   

3.
N‐Isopropyl‐4‐vinylbenzylamine (PVBA) was synthesized and used as an initiator for the polymerization of methacrylates to synthesize macromonomers with terminal styrenic moieties. LiPVBA initiated a living polymerization and block copolymerization of methyl methacrylate, 2‐(N,N‐dimethylamino)ethyl methacrylate and tert‐butyl methacrylate and produced polymers having well‐controlled molecular weights and very low polydispersities (w/n < 1.1) in quantitative yield. 1H NMR analysis revealed that the polymers contained terminal 4‐vinylbenzyl groups. The macromonomers were reactive in the copolymerization with styrene.  相似文献   

4.
The anionic polymerization of methyl methacrylate was carried out in the presence of potassium tert-butoxide (t-BuOK)/quaternary ammonium salts (QAS) in toluene and tetrahydrofuran at −60°C. It was found that in toluene some QAS additives substantially increase the syndiotacticity of poly(methyl methacrylate). Two types of QAS were distinguished, quite different in their action. The addition of QAS with one or two longchain alkyl groups (>C12), does not change significantly the mode of the monomer addition, whereas the polymerization in the presence of tetraalkylammonium salts with four equal substituents and dimethyldidodecylammonium bromide yields predominantly a syndiotactic polymer with high conversion and comparatively low polydispersity (M̄w/M̄w = 1.3−1.5). In some cases QAS additives are more effective modifiers than cryptand [2.2.2].  相似文献   

5.
Polystyrene/poly[styrene-co-(butyl methacrylate)] block copolymers with controlled molecular weights and with polydispersities generally below w/n = 1,45 and partially as low as w/n = 1,19 were synthesized by a free radical bulk copolymerization using a 2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO)-capped polystyrene macroinitiator. The influence of the macroinitiator concentration on the block copolymerization was studied. The polymerization rates are independent of the macroinitiator concentration and are close to that of thermally self-initiated styrene/butyl methacrylate copolymerizations showing the important role of self-initiation for N-oxyl mediated free radical polymerizations.  相似文献   

6.
A combination of tert-butyllithium (t-BuLi) and bis(2,6-di-t-butylphenoxy)methylaluminium (MeAI(ODBP)2) was found to be an efficient initiator for heterotactic living polymerization of certain alkyl methacrylates in toluene at low temperatures. The polymerization of methyl methacrylate (MMA) with t-BuLi/MeAI(ODBP)2 (AI/Li=5 mol/mol) in toluene at −78°C gave heterotactic-rich poly(methyl methacrylate) (PMMA) with narrow molecular weight distributions (MWDs) (heterotactic triad fraction mr = 68%, ratio of weight- to number-average molecular weights M̄w/M̄n = 1.06-1.17). Other alkyl methacrylates also gave heterotactic polymers under the same conditions; in particular, ethyl and butyl methacrylates gave polymers with heterotactic triad fractions of 87%. The highest triad heterotacticity of 91.6% was obtained for the polymerization of ethyl methacrylate at −95°C. Some characteristic features of this stereospecific polymerization were discussed based on the polymerization results combined with other structural information of the polymer such as chain-end stereostructure and stereosequence distribution in the main chain.  相似文献   

7.
A zerovalent nickel complex, Ni(PPh3)4, induced living radical polymerization of methyl methacrylate (MMA) in conjunction with an organic bromide as an initiator [R–Br: CCl3Br, (CH3)2C(CO2Et)Br, (CH3)2C(COPh)Br] in the presence of Al(Oi-Pr)3 additive. The molecular weight distributions were narrow (w/n ∼ 1.2) throughout the reactions, and the number-average molecular weights (n) increased in direct proportion to monomer conversion. In contrast, the polymers obtained with CCl4 in place of R–Br had broader MWDs (w/n > 2). The Al(Oi-Pr)3 additive should be added for the smooth polymerizations of MMA to occur, similarly to those with a divalent nickel bromide, NiBr2(PPh3)2. The Ni(PPh3)4-mediated living polymerization apparently proceeds via the activation of the C Br bond from the initiators R Br, assisted by the redox reaction of the complex between Ni(0) and Ni(I) species. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3003–3009, 1999  相似文献   

8.
The ceric ion-initiated graft copolymerization of methyl methacrylate onto wood cellulose was found to depend on the concentrations of initiator, monomer, and cellulose. The structure of cellulose—methyl methacrylate graft copolymers was studied by hydrolyzing away the cellulose backbone to isolate the grafted poly(methyl methacrylate) branches. The molecular weights and molecular weight distributions of the grafted poly(methyl methacrylate) were determined by using gel-permeation chromatography. The number-average (M?n) molecular weights ranged from 36 000 to 160 000 and the polydispersity ratios (M?w/M?n) varied from 4.0 to 7.0. The grafting frequency or the number of poly(methyl methacrylate) branches per cellulose chain calculated from the per cent grafting and molecular weight data varied from 0.38 to 3.2. The structure of cellulose—methyl methacrylate graft copolymers and the effect of stepwise addition of initiator on the structure are discussed.  相似文献   

9.
Reverse atom transfer radical polymerization of methyl acrylate in the presence of a conventional radical initiator (2,2′-azoisobutyronitrile, AIBN) in bulk was successfully implemented via a new polymerization procedure. The system first reacts at 65–70°C for ten hours, then polymerizes at 100°C. Various mole ratios of AIBN to CuIICl2 were used in this work, all of which result in a well-controlled radical polymerization with high initiation efficiency and narrow molecular weight distribution, i.e., the polydispersity is as low as w/n = 1.36.  相似文献   

10.
Living and highly isotactic poly(methyl methacrylate) (PMMA) anion (M̄n = 2.5 × 103) prepared with t-C4H9MgBr as an initiator was protonated with phenol in toluene at −78°C. The reaction was stereospecific toward meso addition, and the meso/racemo ratio at the chain-end of the resultant polymer was 89/11. Addition of 1,4-dioxane to the living isotactic PMMA anion in toluene at −78°C remarkably reduced the viscosity of the system, and protonation of the PMMA anion with phenol in the presence of 1,4-dioxane enhanced the meso-specificity to 94%. On the other hand, the protonation reaction of the living syndiotactic PMMA anion (M̄n = 2.5 × 103), which was generated by t-C4H9Li/(n-C4H9)3Al in toluene at −93°C, with t-butanol was found to be 97% racemo-specific. These highly stereospecific protonation reactions of the stereoregular PMMA anions were in contrast to the protonation of the anions with methanol or benzyl alcohol which was almost non-stereospecific.  相似文献   

11.
Anionic copolymerizations of butadiene (M1) with excess 1-(4-dimethylaminophenyl)-1-phenylethylene (M2) were conducted in benzene at room temperature for 24–48h using sec-butyllithium as initiator. Anisole, triethylamine and t-butyl methyl ether were added in ratios of [B]/[RLi] = 60, 20, 30, respectively, to promote copolymerization. Narrow molecular weight distribution copolymers with M̄n = 14 × 103 to 32 × 103 g/mol (M̄w/M̄n =1.02–1.03) and 8,12 and 30 amine groups per chain for anisole, triethylamine and t-butyl methyl ether, respectively, were obtained. The butadiene monomer reactivity ratios (r1) were 42, 33 and 14 for anisole, triethylamine and t-butyl methyl ether, respectively.  相似文献   

12.
Glass transition temperatures Tg of methyl methacrylate/butyl acrylate copolymers obtained by means of atom‐transfer radical polymerization are measured using differential scanning calorimetry. Due the nature of this polymerization method an increase in molecular weight is produced as the reaction progresses, which gives rise to an increase in Tg. Simultaneously, a composition gradient with the enrichment of butyl acrylate causes a decrease in Tg. These opposite effects almost compensate each other and, therefore, no influence on the molecular weight at n < 10000 is found. This fact allows the application of the Johnston's equation and the Mayo‐Lewis terminal model to describe and predict the variation in Tg with copolymer conversion for the copolymers and under the experimental conditions investigated.  相似文献   

13.
The use of two kinds of tantalum(V) aminopyridinato complexes, bis(2-benzylaminopyridinato)trichlorotantalum(V) and trichlorobis[2,6-di(phenylamino)pyridinato-N,N′]-tantalum(V), activated by methylaluminoxane was studied in polymerization of ethylene. The activities of these homogeneous catalyst systems are comparable to those of metallocenes. The weight-average molecular weights (w) of the produced polyethylenes are between 60 000 and 200 000 and w/n ≈ 2.  相似文献   

14.
A series of star-branched polyisobutylenes with varying arm molecular weights was synthesized using the 2-chloro-2,4,4-trimethylpentane/TiCl4/pyridine initiating system and divinylbenzene (DVB) as a core-forming comonomer (linking agent). The resulting star-branched polymers were characterized with regard to the weight-average number of arms per star molecule (N̄w) and dilute solution viscosity behavior. As the molecular weight of the arm (M̄w, arm) was increased, dramatically longer star-forming reaction times were needed to produce fully developed star polymers. It was calculated that N̄w varied from 50 to 5 as the M̄w, arm was increased from 13,000 to 54,000 g/mol. The radius of gyration, Rg, of the star polymers was observed to increase as M̄w, arm was increased. The solution properties of the star polymers were evaluated in heptane using dilute solution viscometry. It was determined that the stars had a much higher [η] compared to the respective linear PIB arms, but a much lower [η] compared to a hypothetical linear analog of an equivalent molecular weight. The dependence of [η] on temperature for the stars and linear arms was very small over the temperature range 25 to 75°C, with only a very slight decrease with increasing temperature. [η]star was also determined to increase with increasing M̄w, arm, but decrease with increasing M̄w, star. The branching coefficient, g′, calculated for the stars at 25°C, increased as N̄w decreased and agre ed well with literature values for other star polymer systems. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35: 3767–3778, 1997  相似文献   

15.
The polymerizations of isobutylene initiated with the system tert-butyl chloride (t-BuCl)/SnCl4 and carried out in CH2Cl2 at −20°C and −78°C were investigated. The results obtained demonstrate that the presence of t-BuCl in the polymerizing system gives rise to a PIB product with a distinctly bimodal MWD. The higher-molecular weight (HMW) PIB, n = 20000, I=w/M̄n ∼ 2.5, is the result of existence of the protogenic initiation with residual water in the reaction system. The lower-molecular weight (LMW) PIB, n < 600, w/M̄n ≤ 1.4, is the product of polymerization initiated presumably with a complex t-BuCl-SnCl4-H2O. To elucidate the reaction mechanism of the polymerization initiated with the complex, a series of similar isobutylene polymerizations using the initiation system 2,5-dichloro-2,5-dimethylhexane (DDH)/SnCl4 was run and the oily LMW PIB samples were investigated by 1H-NMR. A new polymerization mechanism describing the role of DDH and t-BuCl is suggested.  相似文献   

16.
Acrylamide was graft polymerized onto the surface of a biodegradable semicrystalline polyester, poly(ε‐caprolactone). Electron beam irradiation at a dose of 5 Mrad was used to generate initiating species in the polyester. The degradation in vitro at pH 7.4 and 37°C in a phosphate buffer solution was studied for untreated, irradiated and acrylamide‐grafted polymers. In the case of poly(ε‐caprolactone), all materials showed similar behavior in terms of weight loss. No significant decrease in weight was observed up to 40 weeks, after which the loss of weight accelerated. The main differences in degradation behavior were found for the average molecular weights, n and w. Virgin poly(ε‐caprolactone) maintained n and w up to about 40 weeks, whereas the irradiated and grafted poly(ε‐caprolactone) showed similar continuous declines in n and w throughout the degradation period. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1651–1657, 1999  相似文献   

17.
Alkyl methacrylate monomers with three kinds of dendritic fragments as alkyl groups, two kinds of Fréchet-type and a new type with 3,5-diphenoxybenzyl as first generation, were prepared by reaction of the corresponding dendritic bromides with methacrylic acid in good yield and were homopolymerized with a radical initiator. Polymers with side groups in first generation had a degree of polymerization (n) of 17–24, markedly higher than those with side groups in second generation, n ≈ 6–7.  相似文献   

18.
Polycondensation of 1,10‐decanediol with dimethyl‐H‐phosphonate taken in excess leads to oligomers with methyl‐H‐phosphonate end groups. The polytransesterification of the resulting oligomer as well as the related model reactions were studied. The synthesis of poly(decamethylene‐H‐phosphonate) was analyzed and the final product had n = 1.4–1.9 104 (from end groups, vpo, and n of the derived polymers). The exchange of the ester groups between two homoesters (dimethyl and diethyl phosphonates) used as models, conducted at r.t. and catalyzed by metal alkoxide provides mixed (hetero) ester in a few minutes. If the concentration of the catalyst is not high enough, then the reaction does not go to equilibrium, because the alcoholate anions are converted into the anions of monoesters of the H‐phosphonic acid, catalytically inactive at this temperature. However, these monoesters become catalytically active at higher temperature, i.e., at the conditions used for preparing higher molecular‐weight products by transesterification. The apparent rate constants () of the ester exchange catalyzed by monoester salt (modeling the propagation step in polytransesterification) were determined by two independent methods; at 130°C ∼ 1.0 · 10−2 mol−1 · L · s−1. The detailed study of the model polytransesterification, and particularly of the polymer end groups appearance and disappearance (studied by 1H‐, 13C‐, and 31P‐NMR) allowed postulation of the reaction mechanism and confirmed our previous work, describing formation at these conditions of polymers with n > 104. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1365–1381, 1999  相似文献   

19.
The chiral heterobimetallic complexes Li[Ln(η5 : η1-C5R41SiMe2NCH2CH2R2)2] (Ln = Y, Lu; C5R41 = C5Me4, C5H4, 3-C5H3 t Bu; R2 = OMe, NMe2; Me: methyl; tBu: tert-butyl) have been found to polymerize ϵ-caprolactone to give a polymer of high molecular weight (n < 20 000) and moderate polydispersity (w/n < 2.0). Failure to observe a correlation between monomer/initiator ratio and molecular weight suggests a polymerization mechanism different from a pseudo-anionic mechanism.  相似文献   

20.
The mechanism of the strictly alternating anionic copolymerization of phenyl glycidyl ether (PGE) and phthalic anhydride (PA) was initiated by various imidazoles. Because of the strictly alternating copolymerization polyesters with a repeating unit of PGE-PA were obtained. The mechanism of the reaction was analyzed by means of matrix-assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF-MS). With this technique the molar masses of the oligomers, the molar mass of the repeating unit, the weight-average molar mass w and the number-average molar mass n, their ratio w/n and the residual molar mass could be calculated. The strictly alternating copolymerization was easy to prove because the molar masses of PGE and PA are slightly different. The question whether the initiator remains chemically bound during the whole reaction could be solved. To this end polyesters obtained by initiation with various imidazoles with different molar masses were synthesized. The calculated residual molar masses correspond exactly to the molar masses of the imidazoles.  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