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 62 毫秒
1.
Radical copolymerization of butyl methacrylate with 2,3-dimethylbutadiene in the presence of Al(C2H5)2Cl or ZnCl2 results in alternating copolymers. The nature of active centers and the mechanism of polymerization in these systems have been studied by means of ESR measurements in combination with calorimetry at low temperatures. The active centers are monoradicals propagating by alternative addition of single monomer molecules; thus the reaction can be described in terms of a conventional kinetic scheme of radical additional polymerization. Participation of binary donor—acceptor complexes of the monomers in the reaction has not been confirmed. Similar conclusions have been drawn for the other alternating system studied, maleic anhydride–2,3-dimethylbutadiene. The feasibility of formation of alternating copolymers in the studied systems by the conventional mechanism of binary radical copolymerization has been confirmed by qualitative quantum-chemical treatment of the propagation reactions with due account to the donor–acceptor interactions in the transition state.  相似文献   

2.
The effect of zinc chloride (ZnCl2) on the cationic polymerization of isobutyl vinyl ether (IBVE) initiated by carboxyl groups on a carbon black surface was investigated. Although the polymerization of IBVE was initiated by carboxyl groups on the surface, the rate of polymerization was small and the molecular weight distribution (MWD) of poly IBVE was very broad. The rate of the polymerization was found to be drastically increased, and 100% monomer conversion was achieved in a short time by the addition of ZnCl2. The number-average molecular weights (Mn) of the polyIBVE were directly proportional to monomer conversion in the polymerization initiated by the carbon black/ZnCl2 system. By addition of the monomer at the end of the first-stage polymerization, the added monomer was smoothly polymerized at the same rate as in the first stage. The Mn of the polymer was in excellent agreement with the calculated value, assuming the polyIBVE chain forms per unit carboxyl group on the surface and MWD was narrow (Mw/Mn = 1.2 ~ 1.3). Based on the results, it is concluded that carbon black/ZnCl2 system has an ability to initiate the living cationic polymerization of IBVE. Furthermore, it was found that polyIBVE was grafted onto the carbon black surface after the quenching of the living polymer with methanol. © 1995 John Wiley & Sons, Inc.  相似文献   

3.
This paper discusses the nature of the living growing species in cationic polymerization from the viewpoint of the steric structure of poly(isobutyl vinyl ether) [poly(IBVE)]. At −78 °C, living polymerization was induced with the HCl-IBVE adduct (1)/ZnCl2 system in a EtNO2/CH2Cl2 mixture, whereas similar systems with EtAlCl2 led to conventional cationic polymerization. In this polar medium, both systems gave polymers with very similar and low isotacticity (meso ≈ 56%), indicating that the propagating reaction is mediated by free ions. Thus, regardless of solvent polarity, or involvement of free ions or ion pairs, living cationic polymerization requires a suitably nucleophilic counteranion. As model reactions of the growing species, 1/ZnCl2 and 1/EtAlCl2 were directly analyzed by 1H NMR spectroscopy.  相似文献   

4.
Polymerization of vinyl acetate (VAc) was carried out with potassium persulfate (KS) and ammonium sulfite as redox initiator in the presence of acetonitrile for the purpose of preparation of polyvinyl alcohol (PVA) of desired degree of polymerization (DP). The PVA obtained by this method at 0°C showed higher DP and lower solubility in water. It was observed that acetonitrile was a useful chain transfer agent in this polymerization system. It was also observed that ZnCl2 accelerated the polymerization of VAc and the over-all activation energy of polymerization was 9.73 Kcal/mol. However, the stereoregularity of PVA obtained by the saponification of PVAc in this system was not improved by the presence of ZnCl2.  相似文献   

5.
The thermal and radiation-induced in-source and postirradiation polymerizations of N-tert-butylacrylamide and (N-tert-butylacrylamide)2–ZnCl2 complex of this monomer were studied at various temperatures. In in-source, solid-state polymerizations of monomer and complex the conversion was about 95% at 21°C in about eight days. Their postirradiation polymerizations were also studied in solid state. The conversion-time curves of these two systems show an autoacceleration as in-source polymerization. In both types of polymerization the overall rate of polymerization of complex was higher than that of pure monomer at the same polymerization temperature. In investigations of the thermal polymerization of N-tert-butylacrylamide and ZnCl2-complex it was observed that the ZnCl2-complex system can be polymerized in air in the molten and solid state. The conversion of monomer to polymer reaches limiting values in solid state in about 1 hr. The thermal polymerization of ZnCl2-complex in the molten state was also studied and 100% conversion was obtained in 30 min. The thermal polymerization of pure monomer was studied in vacuum and an appreciable amount of polymer was obtained in the molten state; however, the thermal polymerization of this monomer is negligible in solid state. In this work rates of polymerization for N-tert-butylacrylamide and (N-tert-butylacrylamide)2–ZnCl2 are compared under various experimental conditions and overall activation energies are calculated.  相似文献   

6.
Abstract

The kinetics of the radical copolymerization of acrylonitrile with methyl acrylate complexed by zinc chloride (ZnCl2) in dimethylformamide (DMF) was investigated at 60, 65, and 70°C. The kinetic data revealed that Rp was an inverse function of ZnCl2 concentration and directly related to monomers concentration. The increase in the activation energy from 11.85 to 19.25 kJ·mol?1 and the decrease in the value of the ratio of the propagation to termination rate constants (kp 2/kt ) from 0.08 to 0.06 L·mol?1·s?1 on the addition of ZnCl2 indicated its retarding effect. The chain transfer constant of DMF for the system was 16.25 × 10?4, accordingly the degree of polymerization decreased. The structure and composition of the copolymers determined by 1H-NMR and elemental analysis was found to be alternating. The nonideal behavior of the glass transition temperatures determined by DSC also favors the alternation of monomer units in the copolymer. The reaction proceeds via a cross-propagation mechanism.  相似文献   

7.
In the present study, the reaction conditions for homodimerization process of 3-acetylcoumarin were achieved under sonication using combination of zinc and metallic salt (ZnCl2 or Zn(OAc)2). Appropriate frequency and sound amplitude have been identified as significant variables for the initiation of the reaction. On the base of first principal calculations and experimental results, the mechanism of the reaction was investigated. The relative stability of the possible intermediates has been compared, including evaluation on the ionic and radical reaction pathways for the dimerization process. Theoretical results suggested that the radical mechanism is more favorable. The C-C bond formation between the calculated radical intermediates occurs spontaneously (∆G = −214 kJ/mol for ZnCl2, −163 kJ/mol in the case of Zn(OAc)2), which proves the possibility for the homodimerization of 3-acetylcoumarin via formation of radical species. Both experimental and theoretical data clarified the activation role of the solvent on the reactivity of the Zn-salt. The formation of complexes of solvent molecules with Zn-atom from the ZnCl2 reduces the energy barrier for the dissociation of Zn-Cl bond and facilitate the formation of the dimeric product.  相似文献   

8.
Abstract

The processes of ZnCl2 selective sorption by macromolecules of polymethyl methacrylate and polybutyl methacrylate in binary mixtures with ZnCl2 have been studied using light scattering and equilibrium dialysis methods. It has been shown that to a certain concentration of ZnCl2, its average concentration in the volume of macro-molecular coils is equal to that of the complexing agent in the solution, i.e., the parameter of the selective sorption β = 0. However, for the system PBMA-BMA-ZnCl2 with increasing [ZnCl2], a negative value of β is observed. After the “threshold” concentration of [ZnCl2]: methyl methacrylate = 0.08, butyl methacrylate = 0.15, an increase in the selective sorption of ZnCl2 concentrations is found. The interaction of ZnCl2 with macromolecules of polymethyl methacrylate and polybutyl methacrylate is accompanied by a change in the steric parameter and in the thermodynamic quality of the solvent. A probable mechanism of ZnCl2 interaction with polymer coils is discussed.

Many reactions involving the participation of macromolecules occur in multicomponent solutions, and the distribution of solvent components, within and outside of a polymer coil, can considerably affect the process features. Hence, the behavior of macromolecule in binary solvents is important in the study of the physical chemistry of polymer solutions. Radical polymerization in the presence of complexing agents (CA), where the latter considerably influence the kinetic features of the reaction [1, 2], is an example of a process in which these effects should be manifested. In particular, in methyl methacrylate (MMA) polymerization an increase in CA (ZnCl2) concentration results in an increase in the chain propagation constant, influences the composition of copolymers, and can cause a sharp drop in the bimolecular termination constant. It has been assumed that the latter is connected with the specific interaction of CA molecules with the propagating macroradicals [3, 4].

This paper reports on studies of the behavior of methacrylic polymers in the binary solvent monomer- ZnCl2. These studies permitted us to simulate the polymerizing system and to learn the effects of complexation on the equilibrium flexibility of a macromolecule. These data are also of independent interest because they permitted us to elucidate the specific characteristics of complexing with the participation of macromolecular ligands.  相似文献   

9.
Monoallyl compounds are not readily homopolymerized by a conventional free‐radical mechanism. However, the polymerization of allylbiguanide hydrochloride was reported to proceed in a concentrated solution of hydrochloric or phosphoric acid in the presence of a radical initiator. Here we have studied the polymerization of allyl alcohol by a radical initiator in the presence of a Lewis acid (ZnCl2, CuCl2 or MgCl2) in an organic solvent (toluene, hexane, methanol or isopropanol). Reactions were performed either at room temperature or 50°C under an atmosphere of nitrogen or in a sealed tube. The same polymerization was also carried out in water and in a concentrated acid solution. The polymer product was purified by dialysis in 0.2–3.7% yield and confirmed by elemental analysis, infrared spectroscopy and 1H NMR. The molecular weight range of poly(allyl alcohol) was 10,000–35,000. The polymerization of allyl acetate by the radical initiator under the above conditions gave poly(allyl acetate) with the molecular weight range of 10,000–13,800 by multi‐angle laser light scattering. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

10.
The reaction between ZnCl2 and N,N-bis[(3,5-dimethyl-1H-pyrazol-1-yl)methyl]-1-phenylethylamine (bdmppea) affords [(bdmppea)ZnCl2], whose structure has been determined by X-ray crystallography. The [(bdmppea)ZnEt2] complex in situ prepared by the reaction between [bdmppea] and ZnEt2 exhibited high activity toward the polymerization reaction of rac-lactide at room temperature. However, its activity decreased sharply with decreasing temperature. Stereospecificity of this catalyst characterized by heterotacticity (Pr) was determined by homonuclear decoupled NMR spectroscopy, which value was ∼0.58.  相似文献   

11.
Polymerization of acrylonitrile photoinitiated by naphthalene, anthracene, phenanthrene, and pyrene is accelerated by an admixture of zinc (II) chloride, acetate, or nitrate. The effect of zinc (II) salts on the rate of pyrene-photoinitiated polymerization of acrylonitrile leads to an increase in this rate in the order Zn/OCOCH3/2 < ZnCl2 < Zn/NO3/2. The maximum polymerization rate is achieved at the molar ratio [ZnCl2]/([ZnCl2] + [pyrene]) approximately 0.7. In contrast to the photoinitiated polymerization of acrylonitrile, the methyl methacrylate admixture of zinc (II) chloride exerts a smaller effect on the polymerization rate. In the pyrene-photoinitiated polymerization of styrene an admixture of zinc (II) chloride retards the polymerization rate. Fluorescence of aromatic hydrocarbon in the system acrylonitrile–aromatic hydrocarbon is efficiently quenched by zinc (II) chloride. Stern–Volmer constants determined for pyrene (80 dm3 mole?1), phenanthrene (66 dm3 mole?1), and naphthalene (49 dm3 mole?1) are higher by about 2–3 orders of the Stern–Volmer constants for fluorescence quenching of aromatic hydrocarbons by acrylonitrile in the absence of ZnCl2. The fluorescence of anthracene in acrylonitrile is not quenched by ZnCl2. The acceleration effect of Zn (II) salts on the polymerization of acrylonitrile photoinitiated by aromatic hydrocarbons depends on two factors: an increase in the ratio of the rate constant of the growth and termination reactions, kp/kt, and an increase in the quenching constant of fluorescence of aromatic hydrocarbon, kq, by the complex {acrylonitrile…ZnCl2}. ZnCl2 thus influences both the growth and initiation reactions of the polymerization process.  相似文献   

12.
Overall rates of polymerization of ring-substituted-N-phenylmaleimides were measured, mainly in dioxane. Rate equations were deduced. The rate of polymerization increased on addition of ZnCl2; the effect was attributed to the interaction of the Lewis acid and the π-cloud of the growing maleimide radical. The effect of solvent on the rate of polymerization was also investigated.  相似文献   

13.
The polymerization kinetics of 4,4'-biphenyldicarbonitrile catalyzed by complex of 4,4'-biphenyl-dicarbonitrile with ZnCl_2 was studied. The cyano group concentration was measured by infrared spectroscopic analysis using potassium ferricyanide as the internal reference. The results proved that the polymerization is a second order reaction, and the activation energy of polymerization is 12.4 kcal/mol. The polymerization rates of 4,4'-biphenyldicarbonitrile catalyzed by other seven complexes were also measured. The polymerization mechanism was discussed.  相似文献   

14.
The cationic polymerization of 2‐[4‐(methoxycarbonyl)phenoxy] ethyl vinyl ether, a vinyl ether with a benzoate pendant, was carried out with an HCl/ZnCl2 initiating system in methylene chloride at −15 °C. The polymerization proceeded with living/long‐lived propagating species to produce polymers with controlled molecular weights and relatively narrow molecular weight distributions (weight‐average molecular weight/number‐average molecular weight ≤ ∼1.4), despite the formation of a small amount of oligomeric products during the polymerization. The structural analysis showed that the lowest molecular weight oligomer had the structure CH3CH(OCH2CH2OC6H4COOCH3)OCH2CH2OC6H4COOCH3. The oligomer was formed by the reaction of the monomeric propagating species with the alcohol produced by the side reaction of the active species with water as an impurity during the early stage of polymerization. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4362–4372, 2000  相似文献   

15.
Phosphoric and phosphinic acid derivatives (R1R2PO2H; R1, R2 = OPh, OPh; OnBu, OnBu; Ph, Ph; Ph, H) in conjunction with zinc chloride (ZnCl2) led to living cationic polymerization of isobutyl vinyl ether (IBVE) in toluene below 0°C. The number-average molecular weights (M?n) of the polymers (M?n > 2 × 104) were directly proportional to monomer conversion and in excellent agreement with the calculated values assuming that one polymer chain forms per R1R2PO2H molecule. Throughout the reaction, the molecular weight distributions (MWDs) stayed narrow (M?w/M?n ? 1.1). A dibasic acid, PhOP (O) (OH)2, coupled with ZnCl2, also induced living cationic polymerization of IBVE where one molecule of the acid generated two living polymer chains. The polymerization by (PhO)2PO2H/ZnCl2 and its model reactions were directly analyzed by 31P and 1H-NMR spectroscopy. The analysis showed that the acid initially forms the adduct [CH3CH(OiBu)OP(O)(OPh)2], the phosphate linkage of which is in turn activated by ZnCl2 so as to initiate living propagation. The finding thus indicates that (PhO)2PO2H indeed acts as an initiator in the living polymerization. The NMR analysis also suggested that an exchange reaction occurs between the phosphate group at the polymer terminal and the chlorine in ZnCl2. The occurrence of living IBVE polymerization with these various R1R2PO2H/ZnCl2 systems shows that phosphoric and phosphinic acids are another general class of protonic acids which are effective initiators for the living cationic polymerization assisted by Lewis acids. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
A novel phosphate monomer, Op‐(methacryloyloxymethyl)benzyl O,O‐diethyl phosphate (MDP) was synthesized by the reaction of diethyl phosphorochloridate with 1,4‐benzenedimethanol, followed by the reaction with methacryloyl chloride in the presence of triethylamine. The radical polymerization of MDP and copolymerization with methyl methacrylate were carried out in the presence of 2,2′‐azobisisobutyronitrile (3 mol %) in dimethylacetamide at 60 °C for 20 h to afford phosphate‐pendant polymers. The polymerization of glycidyl phenyl ether (GPE) was carried out with the phosphate‐pendant polymer as an initiator in the presence of ZnCl2. The polymerization did not proceed below 90 °C but rapidly proceeded above 90 °C to afford polyGPE. The phosphate‐pendant polymer served as a good thermally latent polymeric initiator. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 3365–3370, 2001  相似文献   

17.
On electrochemical initiation of alternating copolymerizations of styrene–acrylonitrile (AN) and styrene–diethyl fumarate (DEF) in the presence of ZnCl2, radical anions of AN–ZnCl2 and DEF–ZnCl2 complexes produced at the cathode were assumed to initiate copolymerization. In analogy with the cathode-initiated copolymerization, the radical anions of AN–ZnCl2 and DEF–ZnCl2, generated with the carbanions such as sodium naphthalene, disodium α-methylstyrene tetramer dianion, and butyllithium, were also found to produce alternating copolymers of styrene–AN and styrene–DEF. On the contrary, no polymers were obtained from methyl methacrylate (MMA)–styrene and methacrylonitrile (MAN)–styrene in the presence of ZnCl2 either with carbanions or by electrochemical reduction. Styrene–MAN–ZnCl2 yielded an alternating copolymer with carbanions upon introduction of oxygen.  相似文献   

18.
Labeled aluminum alcoholate was used to determine the content of end groups of polyethers formed in polymerization reactions initiated by the system Al(OiPr)3 + ZnCl2 (1:1). The polymerization process proceeded in several successive stages. Polymerization initiation and propagation mechanisms were proposed.  相似文献   

19.
Abstract

The initiation mechanism of spontaneous alternating copolymerizations of styrene (St) and some electron-accepting monomers such as methyl methacrylate (MMA), methyl acrylate (MA), methacrylonitrile (MAN), and acrylonitrile (AN) in the presence of ZnCl2 was studied by the spin trapping technique, in which 2-methyl-2-nitrosopropane (BNO) was used as a spin trapping reagent. When this technique was applied to the alternating copolymerization systems of St-MMA-ZnCl2, St-MA-ZnCl2, and St-MAN-ZnCl2, the 2-phenylvinyl radical (·CH[dbnd]CH[sbnd]C6H5) was trapped as nitroxide. The structure of this nitroxide, which showed a large coupling constant (19~20 G) by β-hydrogen, was confirmed by comparison with the result of authentic experiment Accordingly it was concluded that this nitroxide was formed through proton migration from the St cation radical to the acceptor monomer anion radical in the charge- or electron-transfer complex, followed by reaction with BNO.

In the St-AN-ZnCl2 system, however, a nitroxide derived from a cyclic radical was observed together with the nitroxide from 2-phenylvinyl radical. This cyclic radical seemed to be produced via the Diels-Alder adduct between St and AN.  相似文献   

20.
The formation of complexes between ZnCl2 and methyl methacrylate (MMA) or between ZnCl2 and AIBN was tested at two different polymerization temperatures, taking into account the Haeringer and Riess treatment. For any concentration of ZnCl2 between 0.01 and 0.1 mole/liter the formation of ZnCl2?MMA complexes is favored, whereas AIBN?ZnCl2 complexes are hardly showed. The effect of zinc chloride on the stereostructure of poly(methyl methacrylate) was also investigated at both temperatures.  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