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.
Monodispersed poly(4-bromostyrenes) (PBs) and their block copolymers with styrene, isoprene, and 3-methylbutene were prepared and characterized by GPC and NMR. Polystyryl and α-methylstyryl carbanions act as effective initiators of the anionic polymerization of Bs in THF. The undersirable side reactions, due to thermally or photochemically induced decomposition of the bromostyryl carbanions, PBs?, may be eliminated by conducting the reaction at ?78°C and in the dark. Under such conditions, the rate constant of propagation, kp (?78°), is 1.5 × 103 M?1 s?1. Radical anions, \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm Bs}^{\mathop - \limits_ \cdot} $\end{document}, formed as result of electron transfer from sodium naphthalenide to Bs, may eject spontaneously bromine ions. This step and reactions involving the respective phenyl radicals compete at ?78°C with the addition steps leading to polymer formation. Electron affinity of Bs seems to be much higher than that of styrene or isoprene, and PBs? carbanions do not add to the latter monomers. Addition of Bs to polyisoprenyl carbanions leads to formation of the BsIBs block copolymers. BsIBs, prepared in THF, may be converted by preferential hydrogenation of the 1–2 adducts into block copolymers of Bs with poly(2-methylbutene) and isoprenyl segments. The effectiveness of Bs as a flame retarding constituent of polymeric systems seems to be much more pronounced when it is incorporated in a “block” than in a random fashion. A considerably larger fraction of PBs is required to achieve the same LOI value when the respective homopolymers are blended. Spatial distribution of the easily charred microdomains in the block copolymers is believed to be responsible for this phenomenon.  相似文献   

2.
The n-butyllithium-initiated polymerization of a mixture of acrylophenone (AP) and styrene produces only poly(AP), indicating that a chain ending in an AP enolate ion is not sufficiently nucleophilic to add to styrene. Radical copolymerization of AP and styrene yields a polymer containing 65% AP (at 41% conversion). In contrast, lithium dispersion-initiated polymerization of AP and styrene produces a product containing 50–99% AP, depending upon conversion. This observation is discussed in terms of current knowledge concerning alkali metal-initiated polymerization.  相似文献   

3.
A stereospecific synthesis of trans-1-cyanobutadiene is described. This monomer was polymerized by n-butyl lithium in toluene and in tetrahydrofuran to yield sparingly soluble products. In both solvents, the concatenation was predominatly trans-1,4 with a little cis-1,4 and 3,4 content. Copolymers with methyl methacrylate and methyl pentadienoate were prepared in the same way: the reactivity ratios differed markedly in the two solvents being more nearly ideal in tetrahydrofuran.  相似文献   

4.
The present paper discusses the ability of macromonomers to undergo polymerization and copolymerization with acrylic and vinylic monomers. These macromonomers have been synthesized by classical deactivation reactions. Special interest was devoted to macromonomers fitted with polymerizable methylmethacrylate end-groups. The anionic homopolymerization of ω-methacryloyloxy-polystyrene macromonomers was studied in detail and the influence of the molar mass of the macromonomer on the apparent propagation constant was determined. The anionic homopolymerization of ω-methacryloyloxy poly(ethylene oxide) macro-monomers was also examined. In both cases, lithium chloride has to be added in order to reach a better control of the reaction. The dilute solution properties of these polystyrene polymacromonomers have been studied. Some preliminary attempts to apply that anionic homopolymerization of macromonomers to the preparation of “dumbbell” and “palmtree” polymers were presented.  相似文献   

5.
Methacrylaldehyde, methyl methacrylate, methacrylonitrile, styrene and isoprene readily polymerize on potassium–graphite inclusion compounds in ethereal and hydrocarbon solvents. The structure of polymethacrylaldehyde, poly(methyl methacrylate), and polyisoprene as well as the composition of styrene–acrylonitrile and styrene–isoprene copolymers have been investigated. The copolymers have a high content of styrene units which is interpreted in terms of selective adsorption of styrene on the initiator surface.  相似文献   

6.
7.
Anionic copolymerizations of 2,3,4,5,6-pentafluorostyrene (PFS) with 1,3-divinylbenzene (m-DVB) and 1,4-divinylbenzene (p-DVB) were performed by using lithium diisopropylamide as an initiator in order to synthesize the fluorine-containing linear polymer with pendant vinyl groups. The products were soluble copolymers possessing both PFS and DVB monomeric units, and the DVB monomeric unit in copolymer had pendant vinyl group. This copolymerization reaction took a much longer time than that of styrene with DVB. The copolymerization parameter of this system was examined from copolymer composition curves. In this system, m-DVB was found to be more reactive than p-DVB. The reactivity of copolymerization was largely influenced by the reactivity of active species. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Kinetics of anionic copolymerization of styrene (S) and 1,1-diphenylethylene (D) were investigated in THF. The rate constant of addition of D to living polystyrene was found to be k1,2 ± = 250 l./mole-sec. for , Na+ ion-pair, and that for the free , Na+ ion is k1,2?~400,000 l./mole-sec. Both values refer to 25°C. The addition of styrene to ? D?, Na+ was found to be reversible: and k2,1 was determined by three different methods to be ~0.5–0.7 l./mole-sec. Studies performed in a stirred-flow reactor led to k-21 = 13 sec. ?1 and K21 ~ 5 × 10?2 l./mole. An alternating copolymer is obtained in the presence of a large excess of 1,1-diphenylethylene.  相似文献   

9.
It is shown that the products of interaction of ethylene oxide and bicyclic amines containing tertiary nitrogen atoms at the tops of bicyclic structures efficiently initiate the anionic polymerization of acrylonitrile. As opposed to all known initiators of this process, the mentioned initiating systems contain no metal atoms or atoms of elements heavier than oxygen. The polymerization of acrylonitrile under the action of the ethylene oxide–bicyclic amine system in a polar medium (dimethyl sulfoxide) at room temperature occurs in the homogeneous regime over several minutes, while, in a weakly polar solvent (tetrahydrofuran), polymerization occurs in the heterogeneous regime over several hours. The reaction may become homogeneous in a mixture of these solvents at both room temperature and a lower temperature. The number-average molecular masses of the polymers, depending on polymerization conditions, are in the range from 25 × 103 to 480 × 103 and their polydispersity indexes are from 1.55 to ~3.40. It is found that the copolymers of acrylonitrile with oxygen-containing acrylic monomers, as well as with ethylene oxide, can be prepared.  相似文献   

10.
11.
The anionic polymerization of ε-caprolactam in the presence of aromatic polyimides as activators has been studied under adiabatic conditions. It has been shown that despite the crosslinked structure the produced graft copolymers of polycaproamide and polyimides contain the crystalline phase and are characterized by better water resistance and thermal stability, a higher content of the gel fraction, and improved mechanical properties as compared to copolymers prepared under the so-called isothermal conditions. It has been demonstrated that the catalytic system MgBr-ε-caprolactam-aromatic polyimide may be used for the anionic copolymerization of ε-caprolactam with ω-dodecanelactam, and experimental conditions providing a copolymer yield of 75% have been developed.  相似文献   

12.
Anionic copolymerization of p-anisaldehyde (ANA) with dimethylketene (DMK) was made with use of benzophenone–dilithium complex as an initiator at ?78°C in a high vacuum. In spite of a copolymerization in such a good polar solvent as tetrahydrofuran, the composition of the copolymer was nearly exactly 1 : 1 over a quite wide range of the monomer feed. From the analytical data of the product after the hydrogenolysis of the copolymer with lithium aluminum hydride, the copolymer was found to have a structure resulting from the alternating addition of the C?O double bond of ANA to the C?C double bond of DMK. No copolymerizations of ANA with phenyl isocyanate and methyl isocyanate take place under the same conditions.  相似文献   

13.
Diphenyl-, phenylethyl, and phenylmethylketene have been copolymerized with phenyl isocyanate by use of sodium naphthalene in dimethylformamide (DMF) at ?45°C. Reactivity ratios of phenyl isocyanate (r2) with diphenylketene (r1) were r1 = 0.10, r2 = 0.29; with phenylethylketene (r1) were r1 = 1.6, r2 = 0.10; and with phenyl methyl ketene (r1) were r1 = 4.8, r2 = 0.02. The same initiator and solvent system were used for homopolymerization of phenylethylketene and copolymerization with m-chloro-, p-chloro-, p-methoxy-, and m-methoxyphenyl isocyanate as well as with phenyl isocyanate. Molecular weights ranged from 1740 to 4000. The effect of substituents on the order of isocyanate incorporation into the copolymer was m-Cl = p-Cl > m-MeO > H > p-MeO. Phenylethylketene was also copolymerized with m-methoxyphenyl, p-methoxyphenyl, and p-tolyl isocyanate in tetrahydrofuran (THF) at ?78°C. Molecular weights ranged from 2800 to 10,500. The least reactive isocyanate was incorporated into the copolymer to a greater extent in this solvent than in the more polar DMF. DTA showed the presence of crystallinity only in polymers of high isocyanate content. The ketenes copolymerized less readily with alkyl isocyanates, such as ethyl, n-butyl, n-hexyl, and cyclohexyl isocyanate, than with the aromatic isocyanates when sodium naphthalene was used in either DMF or THF.  相似文献   

14.
15.
16.
The copolymerization of isoprene with styrene initiated with sec-butyllithium in cyclohexane solution has been studied by kinetic methods. The rates of the homopolymerization have been measured by normal methods. The rates of the cross-propagation reactions were measured by the rate of appearance or disappearance of the ultraviolet absorption of polystyryllithium when the individual chain ends were reacted with the opposite monomer in the absence of the first. The rates of some of the reactions were also measured in the presence of the other monomer. It was found that polyisoprenyllithium reacted with both monomers with a one quarter-order dependence, and the polystyryllithium with a one half-order dependence. It was found possible to describe the copolymerization in terms of the four individual rate constants and to predict the initial copolymer composition. It was not found necessary to resort to explanations based on preferential absorption of isoprene around the chain ends to explain the high isoprene content of the copolymer.  相似文献   

17.
1,1-Diphenylethylene (M2) and 2,3-dimethylbutadiene (M1) were copolymerized with n-butyllithium in tetrahydrofuran. The rate of consumption of each monomer was followed by the change of high resolution NMR spectra of the reaction mixture. The copolymerization proceeded alternately, if the ratio of initial monomer concentrations, [M2]0/[M1]0, was sufficiently larger than unity. By assuming the rate constant k22 to be zero, the constants k21 were obtained from the consumption rates of the monomers. In the alternating copolymerization, 2,3-dimethylbutadiene was incorporated into the copolymer only as the 1,4-structure, while the 1,2-structure was predominant in homopolymerization.  相似文献   

18.
Block copolymers of the A‐B‐C‐B‐A type were synthesized for the first time via the activated anionic polymerization of hexanelactam (HL) with Na‐HL as an initiator and macroactivators [or polymeric activators (PACs)] as elastificators for nylon‐6. The PACs were prepared by the functionalization of telechelic hydroxyl‐terminated poly(ethylene oxide)–polyisoprene–poly(ethylene oxide) copolymers with different diisocyanates. Hexamethylene diisocyanate (1,6‐diisocyanatohexane) and isophorone diisocyanate (5‐isocyanate1‐isocyanatomethyl‐1,3,3‐trimetylcyclohexane) were used as functionalizing agents. This article reports on the effects that the various central elastomeric PAC blocks (type, content, and molecular weight) had on the polymerization kinetics and on the structure and molecular weights of the multiblock copolymers obtained. The copolymers were characterized spectroscopically. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 487–496, 2003  相似文献   

19.
The butyllithium-initiated polymerization of o- and p-methoxystyrene was studied in toluene at 20°C by dilatometry. Initiation of o-methoxystyrene was found to be instantaneous as evidenced by the absence of any induction period. The propagation rate proceeds by an internal first order with respect to the monomer concentration while the order with respect to the living chain ends varies from 0.67 to 0.51 over a concentration range from 4.5 × 10?4 to 1.8 × 10?2 mole/1. The rate may thus be expressed by the equation, where [M] and [PLi] denote concentration of monomer and poly-o-methoxystyryllithium, respectively, and n varies from 0.67 to 0.51. It is assumed that the propagation proceeds exclusively via the monomeric form of the ion-pairs in analogy with the polymerization of styrene. The variable order results from the relatively high value of the dissociation equilibrium constant of dimeric into the monomeric ion-pairs K that was evaluated graphically to be 10?3 instead of 10?6 for styrene. The propagation rate constant kp was found to be equal to about 50 l./mole-min; the propagation activation energy is equal to 12 kcal/mole. No appreciable termination was found in the polymerization of o-methoxystyrene. On the contrary, no quantitative data could be obtained for the polymerization of p-methyoxystyrene due to a slow initiation and a relatively fast termination reaction with formation of a precipitate of highly branched or crosslinked polymer. It is assumed that this precipitate results from a secondary ring metallation reaction.  相似文献   

20.
Anionic polymerization of vinyl chloride has been studied. Of the organometallic compounds tested as initiators, only butyllithium was found to initiate polymerization. Polymerization in bulk at 0°C and with tert-butyllithium as initiator gave poly(vinyl chloride) in a yield of 38% with M n = 55,000. Tacticity of the anionic PVC was similar to that of conventional PVC prepared at similar temperatures. Anionic PVC was found to be less branched and more heat-stable than the conventional polymer.  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