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.
2.
A study of the radical copolymerization of vinyl chloride (C) and glycidyl acrylate (A) at 60° in dichloroethane solution leads to the following reactivity ratios rc = 0·14 ± 0·02 and rA = 7·4 ± 0·3According to the nature of the solvent, the reaction of protonic acids may cause either cationic polymerization of epoxy groups or addition of HCl or water onto the same groups. The HCl addition is observed during the thermal degradation of the copolymers so that HCl evolution is greatly delayed.  相似文献   

3.
Copolymerization of vinyl cyclohexane and α-methyl vinyl cyclohexane with acrylonitrile in the presence of a complexing agent AlEtCl2 results in the formation of alternate copolymers. In the copolymerization of vinyl cyclohexane with acrylonitrile the copolymer composition depends on the ratio of acrylonitrile to AlEtCl2. If this ratio is unity, alternating copolymers of the composition 1:1 are formed; with a ratio greater than unity statistical copolymers that contain more than 50% acrylonitrile units are produced. The 1H-NMR spectroscopy measurements indicate that the interaction between the comonomers and the complexing agent leads to the formation of ternary donor–acceptor complexes of equimolar composition. The equilibrium constants of these complexes at ?60°C have been determined. The effects of temperature, nature of solvent and dilution on the yield, and composition of the copolymers of vinyl cyclohexane with acrylonitrile formed have been studied. By lowering the temperature the yield of copolymers increases but their composition remains equimolar. An increase in the polarity of the medium results in an increase in copolymer yield, whereas the yield decreases if the reaction is conducted in a donor-solvent medium. Dilution of the reaction mixture disrupts the alternation of units in the macrochain of copolymers. The kinetic pecularities of copolymerization have been investigated. The linear dependence of the copolymerization rate on the product of comonomer concentration is observed. The rate of copolymerization is proportional to the square root of the incident light intensity. Various additions of radical type and irradiation accelerate the process of copolymerization. The mechanism of alternating copolymerization of vinyl cyclohexane monomers with acrylonitrile in the presence of AlEtCl2 is discussed in terms of homopolymerization of the comonomer complex.  相似文献   

4.
The radical polymerization of N-(2-hydroxypropyl)methacrylamide was investigated kinetically. The hydrophilic character of the polymerization medium was found to affect the rate of decomposition of the initiator [2,2′-azobis(methyl isobutyrate)] and the course of primary radical termination. The presence of the -OH group in the alkyl group attached to the nitrogen atom leads to an increase in the molecular weight of the polymer in comparison with polymers of N-alkyl methacrylamides. This phenomenon was interpreted in terms of the possibility of a polymeranalogous transesteramidation and of an increased possibility of transfer to monomer and polymer. The copolymerization parameters of N-(2-hydroxypropyl)methacrylamide (M1) with methyl methacrylate and styrene were determined; in the first case, r1 = 0·84 ± 0·05, r2 = 0·66 ± 0·07; in the second case, r1 = 0·53 ± 0·08, r2 = 1·72 ± 0·19.  相似文献   

5.
Homopolymerization of α-olefins (1-CnH2n, n = 6, 8, 10, 12, 16 and 18) and their copolymerization with styrene were carried out in toluene at 60 °C using diphenylzinc-ethenylbisindenylzirconium dichloride-methylaluminoxane as initiator system. Atactic polystyrene and almost isotactic poly(α-olefin)s were obtained. Copolymerization of S/α-olefin with this initiator system gave isotactic olefin-enriched copolymers. According to DSC analysis, the homopolymers P(1-C12H24), P(1-C16H32), and P(1-C18H36) as well their styrene copolymers are crystalline.  相似文献   

6.
N-(2-thiazolyl)methacrylamide (TMA) monomer was synthesized from 2-aminothiazole by two different methods. The homo- and copolymerization of this monomer with methyl methacrylate (MMA), styrene (St), acrylonitrile (AN), and vinyl acetate (VA) were performed in dimethyl formamide using 1 mol% AIBN at 70°C. The copolymerization behavior was studied in a wide composition interval with the mole fractions of TMA ranging from 0.1 to 0.7 in the feed. Characterization using FTIR and 1HNMR techniques confirmed the structure of the monomer and the prepared homo- and copolymers, but the copolymers compositions were determined from sulphur analysis. The monomer reactivity ratios were computed using Fineman and Ross and Kelen and Tüdös methods for the systems TMA-MMA, TMA-St, TMA-AN and TMA-VA and were found to be r 1 = 0.59 ± 0.05, r 2 = 2.72 ± 0.03; r 1 = 0.39 ± 0.02, r 2 = 0.90 ± 0.01; r 1 = 0.77 ± 0.06, r 2 = 1.99 ± 0.04 and r 1 = 0.80 ± 0.08, r 2 = 0.40 ± 0.05 respectively (r 1 corresponds to monomer reactivity ratio of TMA). The Q and e values for TMA monomer were found to be 1.079 and ?0.054. The synthesized monomer and polymers were tested in vitro for biological activity against some microorganisms, using the disk diffusion technique. Generally, all the polymers were effective against the tested microorganisms, but their growth-inhibition effects varied.  相似文献   

7.
New copolymers of the vinyl saccharide 2-deoxy-2-methacrylamido-D-glucose (M1) with acrylic and methacrylic (M2) acids differing in composition and molecular mass have been synthesized by free-radical copolymerization. The relative activities of the comonomers are determined. It is found that, for acrylic acid, r 1 = 3.03 ± 0.15 and r 2 = 0.5 ± 0.08 and, for methacrylic acid, r 1 = 1.070 ± 0.1 and r 2 = 1.18 ± 0.13. As is evidenced by potentiometric and viscometric measurements, the vinyl saccharide and acid units are capable of interacting, a circumstance that affects the conformational states of macromolecules.  相似文献   

8.
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   

9.
The synthesis of [1‐(fluoromethyl)vinyl]benzene (or α‐(fluoromethyl)styrene, FMB) and its radical copolymerization with chlorotrifluorethylene (CTFE), initiated by tert‐butyl peroxypivalate (TBPPi) are presented. The allyl monomer [H2C = C(CH2F)C6H5] was obtained by electrophilic fluorodesilylation of trimethyl(2‐phenylprop‐2‐en‐1‐yl)silane in 93% yield. A series of seven copolymerization reactions were carried out starting from initial [CTFE]0/([FMB]0 + [CTFE]0) molar ratios ranging from 19.6 to 90.0 mol %. The molar compositions of the obtained poly(CTFE‐co‐FMB) copolymers were assessed by means of 19F nuclear magnetic resonance spectroscopy. Statistic copolymers were produced with molar masses ranging between 13,800 and 25,600 g/mol. From the Kelen and Tudos method, the kinetics of the copolymerization led to the determination of the reactivity ratios, ri, of both comonomers (rCTFE = 0.4 ± 0.2 and rFMB = 3.7 ± 1.8 at 74 °C) showing that FMB is more reactive than CTFE as well as other halogenated or nonhalogenated monomers involved in the radical copolymerization with CTFE. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 3843–3850, 2007  相似文献   

10.
Allyl allyl sulfonate (AAS) has been polymerized under the influence of azobisisobutyronitrile to low molecular weight polymers containing cyclic structures. This is in contrast to the behavior of allyl ethane sulfonate (AES) and of propyl allyl sulfonate (PAS) which did not polymerize under the same conditions. AAS has been copolymerized with styrene, methyl acrylate, and vinyl acetate. The following copolymerization reactivity factors have been found:

rAAS 0.01 ± 0.01 rstyrene 13 ± 1

rAAS 0-37 ± 0.09 rmethyl acrylate 5.3 ± 0-7

rAAS 1.54 ± 0.08 rvinyl acetate 0.5 ± 0.15

The results indicate that AAS has a higher reactivity than AES or PAS.  相似文献   

11.
For studying the photochemistry of carbonyl chromophores in the side-chain, methacrylic esters of para-acylated 2-phenoxyethanols (CH2 = C(CH3) · CO · O · CH2 · CH2O · C6H4 · CO · R), soluble copolymers with styrene and soluble homopolymers were prepared. Comparison of low temperature emission spectra of model compounds, homopolymers and copolymers doped in polystyrene film indicated some interaction between the excited and the ground state structural units in homopolymers. Quantum yield of main chain scission of copolymers of styrene with monomers 1–3 (R = CH3, C2H5, C6H5) at 313 nm radiation in benzene were about 10?4; the cross-linking was the main reaction for copolymer styrene/monomer 4 (R = C6H5CH2). On exposure of copolymers styrene/monomers 1–4 and polystyrene doped with model compounds in film to 313 nm radiation in air, accelerated photo-oxidation occurs as well as cross-linking. Only chromophores of monomers 3 and 4 were effective as sensitizers of photochemical addition of maleic anhydride to benzene by radiation with γ > 340 nm. The difference in the efficiency between model compounds and copolymers on the one hand and a homopolymer on the other hand is due to self-quenching.  相似文献   

12.
(Vinyl acetate)/(ethyl acrylate) (V/E) and (vinyl acetate)/(butyl acrylate) (V/B) copolymers were prepared by free radical solution polymerization. 1H-NMR spectra of copolymers were used for calculation of copolymer composition. The copolymer composition data were used for determining reactivity ratios for the copolymerization of vinyl acetate with ethyl acrylate and butyl acrylate by Kelen-Tudos (KT) and nonlinear Error in Variables methods (EVM). The reactivity ratios obtained are rv = 0.03 ± 0.03, rE = 4.68 ± 1.70 (KT method); rv = 0.03 ± 0.01, rE = 4.60 ± 0.65 (EV method) for (V/E) copolymers and rv ? 0.03 ± 0.01, rB ? 6.67 ± 2.17 (KT method); rv = 0.03 ± 0.01, rB = 7.43 ± 0.71 (EV method) for (V/B) copolymers. Microstructure was obtained in terms of the distribution of V- and E-centered triads and V- and B-centered triads for (V/E) and (V/B) copolymers respectively. Homonuclear 1H 2D-COSY NMR spectra were also recorded to ascertain the existence of coupling between protons in (V/E) as well as (V/B) copolymers. © 1995 John Wiley & Sons, Inc.  相似文献   

13.
Copolymers are formed between styrene and 1-alkyl-4,5-dicyano-2-vinylimidazoles (1-alkylvinazenes). The reactivity ratios for the 1-ethylvinazenes are determined to be high, but also appear to vary with feed composition, with low feed ratios of 1-ethylvinazene giving rEtVz=35 ± 6 and rSty=0.077 ± 0.029, whereas at higher 1-ethylvinazene feed fractions, rEtVz=2.5 ± 2 and rSty=0.089±0.021. Mediation of the 1-methylvinazene/styrene copolymerization with 2,2,6,6-tetramethylpiperidine-N-oxide (TEMPO) leads to a more controlled polymerization which moves closer to ideal copolymerization behavior. The copolymers form with polydispersities approaching 1.5, suggesting bimolecular termination kinetics. The copolymers comply with the Fox relation for Tg.  相似文献   

14.
Autoionizing Rydberg series of Li2 have been observed in the two-step optical cxcitation of a supersonic lithium beam. The series limits are vibrational states of Li2+. In the most probable assignment IP(Li2) = 41236.4 ± 2.5 cm?1 and for Li2+ωe = 263.45 ± 1.3 cm?1; ωeχe = 1.35 ± O.2 cm?1; re = 3.032 ± 0.01 Å; De = 10807 ± 150 cm?1.  相似文献   

15.
Aminomethylated derivatives 2-propenyl-, 2-allyl-, and 4-isopropenylphenols are studied as comonomers in free-radical copolymerization with styrene in bulk. As evidenced by NMR studies, in o-aminomethyl derivatives of alkenylphenols, stable intramolecular hydrogen bonds form between nitrogen atom of the aminomethyl fragment and hydrogen atom of the hydroxyl group; these bonds are not destroyed up to 80°C Therefore, these compounds may be involved in free-radical copolymerization with styrene in which the inhibiting effect of OH groups is avoided at a temperature of 60°C. The values of r 1 and r 2 are determined for two pairs of comonomers: 6-morpholinomethyl-2-propenylphenol and styrene (r 1 = 0.20 ± 0.01 and r 2 = 2.40 ± 0.04) and 2-allyl-6-morpholinomethylphenol and styrene (r 1 = 0.090 ± 0.009 and r 2 = 2.40 ± 0.04).  相似文献   

16.
Apparent transfer constants have been determined for styrene, methyl methacrylate vinyl acetate, and diethyl maleate polymerized in N-allylstearamide at 90°C. Regression coefficients for transfer were: methyl methacrylate, 0.301 × 10?3; styrene, with no added initiator, 0.582 × 10?3; styrene, initiated with benzoyl peroxide, 0.830 × 10?3; vinyl acetate, 62.01 × 10?3; and diethyl maleate, 2.24 × 10?3. Rates of polymerization were retarded for both styrene and methyl methacrylate. Vinyl monomer and comonomer disappearance followed an increasing exponential dependence on both initiator and monomer concentration. Although degradative chain transfer probably caused most of the retardation, the cross-termination effect was not eliminated as a contribution factor. Rates for the vinyl acetate copolymerization were somewhat retarded, even though initiator consumption was large because of induced decomposition. The kinetic and transfer data indicated that the reactive monomers added radicals readily, but that rates were lowered by degradative chain transfer. Growing chains were terminated at only moderate rates of transfer. Unreactive monomers added radicals less easily, producing reactive radicals, which transferred rapidly, so that molecular weights were lowered precipitously. Although induced initiator decomposition occurred, rates were still retarded by degradative chain transfer. A simple empirical relation was found between the reciprocal number-average degree of polymerization, 1/X?n1 and the mole fraction of allylic comonomer entering the copolymer F2, which permitted estimation of the molecular weight of copolymers of vinyl monomers with allylic comonomers. This equation should be applicable when monomer transfer constants for each homopolymer are known and when osmometric molecular weights of one or two copolymers of low allylic content have been determined.  相似文献   

17.
A Fortran IV program for determining copolymerization reactivity ratios is proposed. The program is based on the curve-fitting method and has the advantage of delivering values free of personal judgement. To check its validity the system benzylacrylate (BeA)/methylmethacrylate (MMA) was investigated. The reactivity ratios obtained from the Fineman-Ross plots (r1 = 0 · 34 ± 10 per cent and r2 = 1 · 7 ± 10 per cent) are in good agreement with values obtained by using the proposed method (r1 = 0 · 36 and r2 = 1 · 78).  相似文献   

18.
The radical copolymerization of limonene with styrene by azobisisobutyronitrile in xylene at 80 ± 0.1 °C for 2 h, under inert atmosphere of N2, yields alternating copolymers. The kinetic expression is Rp∝[I]0.5[Sty]1.0[Lim]−1.0. The overall activation energy is calculated as 41 kJ/mol. The FTIR and 1H-NMR spectra of copolymers show bands at 3000 and 1715 cm−1 and peaks at 6.8 δ and 5.3 δ due to phenyl protons of styrene and trisubstituted olefinic protons of limonene, respectively. The values of reactivity ratios r1(Sty)=0.0625 and r2(Lim)=0.014, calculated by Kelen-Tüdos method. The Alfrey-Price Q-e parameters for limonene are 0.438 and −0.748, respectively. The penultimate unit effect is favoured in the present system and the value of φ is 38.49.  相似文献   

19.
The formation of VC-SO2 and VC-(SO2)2 complexes in liquid mixtures of vinyl chloride (VC) and sulphur dioxide has been shown by (a) the freezing point composition diagram and (b) chemical shifts in the PMR spectrum of VC over the complete composition range. It is postulated that SO2 can associate with the CC bond and the Cl atom. These complexes may be involved in the copolymerization and influence the composition and stereochemistry of the product. PMR spectra of VC-SO2-ethane(E) mixtures with [SO2] ? [E] ? [VC] gave Kv = 2·0 ± 0·5, 1·5 ± 0·1 and 1·1 ± 0·3 at 232·6, 272·6 and 301·3 K with ΔHf0Hf = ?6·6 ± 1·4 kJ mol?1 for the VC -(SO2)2 complex. The chemical shift of the trans β-proton was twice that of the other two protons. indicating that SO2 adopts an asymmetric orientation to the double bond.  相似文献   

20.
Combined systems including diphenylzinc (Ph2Zn), a metallocene, and methylaluminoxane (MAO), have been employed to initiate the copolymerization of styrene (S) with p-alkylsubstituted styrenes and with α-olefins. The copolymerization processes depend largely on the comonomer, the nature of the metallocene included in the initiator system, the presence of Ph2Zn, the polymerization temperature and the solvent used. Titanocenes produced true copolymers for S/p-substituted styrene, but not in S/α-olefin copolymerization. On the other hand zirconocenes either did not copolymerize S/p-substituted styrene or produced very low conversions, while they succeeded in copolymerizing S/α-olefin, depending on the particular zirconocene employed. A low p-methylstyrene (p-MeS) content in the S/p-MeS copolymer and a low p-tertbutylstyrene (p-ButS) content in the S/p-ButS copolymer decreased Tm, making them easier to process material than s-PS.  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