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 984 毫秒
1.
Summary The copolymerization of N,N-diethylacrylamide (Ml) with methyl acrylate (M2) was investigated and reactivity ratiosr 1= 0.41 andr 2 = 0.52 obtained. Also the distribution of diad fractions was calculated and the results were interpreted in terms of the product of reactivity ratios. The tendency of the two monomers to alternate was explained on the basis of differences in polatities between the double bonds, this explanation being supported both by the values ofe parameter and NMR spectroscopy data. A copolymerization mechanism was suggested.With 5 figures and 2 tables  相似文献   

2.
Copolymers of the cyclic ketene acetals, 2-methylene-5,5-dimethyl-1,3-dioxane, 3 , (M1) with 2-methylene-1,3-dioxolane, 4 , (M2) or 2-methylene-1,3-dioxane, 5 , (M2), were synthesized by cationic copolymerization. An experimental method was designed to study the reactivity of these very reactive and extremely acid sensitive cyclic ketene acetal monomers. The reactivity ratios, calculated using a computer program based on a nonlinear minimization algorithm, were r1 = 6.36 and r2 = 1.25 for the copolymerization of 3 with 4 , and r1 = 1.56 and r2 = 1.42 for the copolymerization of 3 with 5. FTIR and 1H-NMR spectra when combined with the values of r1 and r2 showed that these copolymers were formed by a cationic 1,2-polymerization (ring-retained) route. Furthermore the tendency existed to form very short blocks of M1 or M2 within the copolymers. Cationic copolymerization of cyclic ketene acetals have the potential to be used for synthesis of novel polymers. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
Abstract

A simple procedure using triangular coordinates for representing triad concentrations as a function of terpolymer compositions (and monomer proportions) is presented. Equal triad concentrations are represented by concentric closed loops or rings of equal triad concentrations converging to a unique point of highest triad concentration. The technique is illustrated with several common terpolymer systems. Alternation in terpolymer systems is assessed by determination of heterotriad concentrations. These results are compared with alternation in component binary systems. An equation is derived for calculating P 12 P MAX 21, maximum 1,2-dyad concentration (maximum mol fraction alternation) from r 1 r 2 product at equimolar copolymer. Alternatively, an equation is proposed for calculating P 12 P MAX 21 from Q-e values. Uses and limitations of r 1 r 2 product in assessing alternation in binary copolymers are discussed.  相似文献   

4.
The static mechanical (uniaxial compression at a low strain rate $\dot \varepsilon $ ≤ 10?4 s?1) and thermomechanical properties of triethylene glycol dimethacrylate (M1) copolymers with vinyl monomers (M2) were studied and analyzed by means of a program developed earlier for identification of structural and physical features of macromolecular networks. The copolymer composition for each monomer pair was varied over the entire range of 0–100 mol % with a step size of 25 mol %. As M2, monomers that ensure different types of unit distribution in chains were selected: alkyl methacrylates (random distribution), styrene (unit alternation tendency), butyl acrylate (block formation tendency), and vinyl acetate (tendency to form chains that are enriched initially in M1 and finally, as M1 is consumed, in M2). To reveal the possible shielding effect capable of lowering the likelihood of the side process of intrachain crosslinking (cyclization), alkyl methacrylates with different degrees of bulkiness of the alkyl radical—methyl, butyl, and lauryl—were used. A systematic study of this representative set of test compounds showed that copolymerization imparted unusual and a priori unpredictable properties to the copolymers formed.  相似文献   

5.
The homopolymerization and copolymerization of butadiene-1-carboxylic acid (Bu-1-Acid) (M1) were studied in tetrahydrofuran at 50°C with azobisisobutyronitrile as an initiator. The initial rate of polymerization was proportional to [AIBN]1/2 and [Bu-1-Acid]1. The overall activation energy for the polymerization was 22.87 kcal/mole. For copolymerization with styrene (M2) and acrylonitrile (M2), the monomer reactivity ratios r1, r2 were determined by the Fineman-Ross method, as follows; r1 = 5.55, r2 = 0.08 (M2 = styrene); r1 = 11.0, r2 = 0.03 (M2 = acrylonitrile). Alfrey-Price Q-e values calculated from these values were 6.0 and +0.11, respectively. The Bu-1-Acid unit in the copolymer as well as the homopolymer was found from infrared and NMR spectral analyses to be composed of a trans-1,4 bond. The hydrogen-transfer polymerization of Bu-1-Acid leading to polyester was attempted with triphenylphosphine as initiator, but did not occur.  相似文献   

6.
An absolute value of kr of ethyl radicals at 860 ± 17°K of 4.5 × 109 M?1·sec?1 was determined under VLPP conditions, where the value of kr/kr should be about 1/2. Thus kr(M?1·sec?1) ~ 1010 at 860°K. An error of as much as a factor of 2 in kr would be surprising, but possible. The value of 1010M?1·sec?1 seems to be a factor of from 2 to 5 too high to be compatible with extensive data on the reverse reaction and the accepted thermochemistry. Changes in the heat of formation and entropy of the ethyl radical can change the situation somewhat, but even these changes when applied to the work of Hiatt and Benson [3] indicate that ethyl combination should be ~ 109.3 M?1·sec?1. More work is necessary if a better value is desired.  相似文献   

7.
Short-range interactions between chain units of random copolymers in solution may be influenced by the composition or precisely by the distribution of sequence lengths of the same monomer units. Steric factors were derived for random copolymers of styrene and acrylonitrile with different compositions from the relation between the limiting viscosity number and the molecular weight. Mark-Houwink relations were obtained in methyl ethyl ketone (MEK) or in N,N′-dimethylformamide (DMF) at 30°C. for random copolymers containing 0.383 (Co-1) and 0.626 (Co-2) mole fraction of acrylonitrile, the expressions are: [η] = 3.6 X 10?4 M w0.62, for Co-1 in MEK; [η] = 5.3 X 10?4 M w0.61, for Co-2 in MEK; [η] = 1.2 × 10?4M w0.77 for Co-2 in DMF. With the Stockmayer-Fixman expression, these correlations become, respectively: [η]/M1/2 = 1.24 × 10?3 + 8.0 × 10?7 M1/2; and [η]/M1/2 = 1.70 × 10?3 + 6.3 × 10?7 M1/2; and [η]/M1/2 = 1.68 × 10?3 + 31.3 × 10?7 M1/2. From the unperturbed mean-square end-to-end distances, 〈L20, determined from the first terms of the latter expressions, together with 〈L20f calculated by assuming the completely free rotation, gives the steric factor σ = (〈L20/〈L20f)1/2 as 2.25 ± 0.05 for Co-1, and 2.31 ± 0.10 for Co-2. These values of σ are close to those for polystyrene (σ = 2.22 ± 0.05) and for polyacrylonitrile (σ = 2.20 ± 0.05). Therefore, it is concluded that the dimensions of random copolymers of styrene and acrylonitrile in solution are not significantly influenced by the composition. In other words, the unperturbed dimensions are not affected by a change in the alternation tendency between styrene units with phenyl side groups having a large molar volume and acrylonitrile units with nitrile groups responsible for the electrostatic interactions. On the other hand, the long-range interactions reflect the effect of sequence length. The Huggins constant and the second virial coefficient obtained from the light-scattering measurements have optimum values at about 0.5 mole fraction of acrylonitrile, where the greatest tendency for alternation seems to exist.  相似文献   

8.
2-Isopropenyl-4-isopropyl-2-oxazolin-5-one (M2), was copolymerized with styrene (M1), and the monomer reactivity ratios were determined to be r1 = 0.31 ± 0.03, r2 = 1.12 ± 0.10. New isomerized oxazolones (M2), 2-isopropylidene-4-methyl-3-oxazolin-5-one, 2-isopropylidene-4-isopropyl-3-oxazolin-5-one, and 2-isopropylidene-4-isobutyl-3-oxazolin-5-one were prepared and copolymerized with styrene. The monomer reactivity ratios were: r1 = 0.36 = 0.07, r2 = 0.0; r1 = 0.39 ± 0.06, r2 = 0.00 ± 0.10; r1 = 0.39 ± 0.10, r2 = 0.0, respectively. The isomerized oxazolones showed no tendency towards homopolymerization by radical initiator. From the results of infrared and NMR spectra and hydrolysis of the copolymer, it was indicated that the isomerized oxazolones participated in copolymerization in the form of 1–4 polymerization of the conjugated dienes (exo double bond at C2 and the C?N in the ring). Copolymers reacted with nucleophilic reagents such as amines and alcohols.  相似文献   

9.
采用新型氨基凝胶自燃法成功制备出尖晶石结构MFe2O4(M=Ca,Mg,Cu,Zn)纳米晶粉末。对合成粉体样品的物相、形貌和磁性能进行了详细的研究。经能量色散X射线谱分析确定了合成MFe2O4粉末的高纯度。系统地研究了所合成的MFe2O4纳米晶粉末的磁性能。所有样品的磁滞回线均较窄,表明了它们具有软磁的特征。经测试得出4种铁氧体的饱和磁化强度(Ms)分别为2.1,29.3,24.1和4.2 emu·g-1;剩余磁化强度(Mr)分别为0.2,2.3,11.4和0.2 emu·g-1。这4种铁氧体样品的Mr/Ms值均小于0.5。对CaFe2O4和MgFe2O4两种典型铁氧体的零场冷却和场冷磁性能作了详细的研究。其中CaFe2O4样品的磁化强度在75 K以下有不一致的变化趋势,这是由于其发生了磁相变。  相似文献   

10.
The monomer reactivity ratios for copolymerization of 2-vinyl-4,4-dimethylazlactone (VA) and ethyl α-hydroxymethylacrylate (EHMA) were 0.20–0.24 and 0.53–0.74, respectively, which show that EHMA is slightly more reactive with VA than with itself and should lead to random copolymers favoring alternation. The VA–styrene (VA–St) system also has a tendency to form random copolymers but with increased tendency for alternation with both r1 and r2 between 0.18–0.22. Tg's of VA–EHMA and VA–St copolymers varied between 100 and 136°C, and 96 and 117°C, respectively. Thermolysis of VA–EHMA copolymers resulted in crosslinking via the ring-opening reaction of VA groups by EHMA alcohols, followed by transesterification involving EHMA units at higher temperatures leading to highly crosslinked structures. The performed dimer of EHMA and VA was also synthesized and found to be an effective crosslinking agent in free radical vinyl polymerizations.  相似文献   

11.
The water-soluble monomers, 1-methyl-4-vinylimidazole, 1-methyl-5-vinylimidazole, 1-ethyl-5-vinylimidazole, and 1-propyl-5-vinylimidazole have been synthesized, polymerized, and copolymerized with 4(5)-vinylimidazole. The copolymers were characterized by 14C-labeling, NMR, pKa determination and viscosity measurements. The monomer reactivity ratios determined by 14C counting are r1 = 1.04; r2 = 0.94 [M1 = 4(5)-vinylimidazole, M2 = 1-methyl-4-vinylimidazole] and r1 = 1.01; r2 = 0.86 [M1 = 4(5)-vinylimidazole, M2 = 1-methyl-5-vinylimidazole]. The esterolytic activity of the copolymers for the hydrolysis of p-nitrophenyl acetate (PNPA) at pH 7–8 in 28.5% ethanol–water was higher than that of the mixtures of homopolymers. At pH 5–6 the esterolytic activities of the copolymers and the mixtures were similar. The most efficient esterolytic activity for PNPA hydrolysis at pH 7.11 in 28.5% ethanol–water occurred for copolymers containing 75 mole % 4(5)-vinylimidazole and for copolymers containing 1-methyl-4-vinylimidazole rather than 1-methyl-5-vinylimidazole.  相似文献   

12.
Two tetranuclear clusters of formula [M4L4(HOMe)4] {H2L = (E)‐1‐[(2‐(hydroxymethyl)phenylimino)methyl]naphthalen‐2‐ol} [M = Co ( 1 ), Ni ( 2 )] were hydrothermally synthesized by reaction of M(OAc)2 · 4H2O with H2L and NaOH in MeOH. X‐ray crystal structure analysis revealed that complexes 1 and 2 are isostructural. In the core of the structures, four MII ions and four oxygen atoms occupied alternate vertices of [M4O4] cubane. The magnetic property measurements of 1 and 2 revealed that overall ferromagnetic MII ··· MII exchange interactions exist in both complexes.  相似文献   

13.
Copolymers of 2-sulfoethyl methacrylate, (SEM) were prepared with ethyl methacrylate, ethyl acrylate, vinylidene chloride, and styrene in 1,2-dimethoxyethane solution with N,N′-azobisisobutyronitrile as initiator. The monomer reactivity ratios with SEM (M1) were: vinylidene chloride, r1 = 3.6 ± 0.5, r2 = 0.22 ± 0.03; ethyl acrylate, r1 = 3.2 ± 0.6, r2 = 0.30 ± 0.05; ethyl methacrylate, r1 = 2.0 ± 0.4, r2 = 1.0 ± 0.1; styrene, r1 = 0.6 ± 0.2, r2 = 0.37 ± 0.03. The values of the copolymerization parameters calculated from the monomer reactivity ratios were e = +0.6 and Q = 1.4. Comparison of the monomer reactivities indicates that SEM is similar to ethyl methacrylate with regard to copolymerization reactivity in 1,2-dimethoxyethane solution. The sodium salt of 2-sulfoethyl methacrylate, SEM?Na, was copolymerized with 2-hydroxyethyl methacrylate (M2) in water solution. Reactivity ratios of r1 = 0.7 ± 0.1 and r2 = 1.6 ± 0.1 were obtained, indicating a lower reactivity of SEM?Na in water as compared to SEM in 1,2-dimethoxyethane. This decreased reactivity was attributed to greater ionic repulsion between reacting species in the aqueous medium.  相似文献   

14.
A novel copolymer of vinylidene cyanide (VCN) and 2,2,2‐trifluoroethyl methacrylate (MATRIF) was synthesized by bulk free radical process in a 52% yield from an equimolar comonomer feed. The copolymer's composition and microstructure were analyzed by FTIR, 1H‐ and 13C‐NMR spectroscopy, SEC, and elemental analysis. The reactivity ratios calculated from both the Q‐e Alfrey‐Price parameters and the Jenkins' Patterns Scheme indicate a tendency to alternation in the copolymerization, the latter method suggesting that MATRIF homopropagation be slightly favoured (rV = r12 = 0.1, rM = r21 = 0.3). The molar incorporation of VCN in the copolymer was only 42 mol % according to the 9.0 wt % nitrogen content determined by elemental analysis, in good agreement with the value obtained by 1H‐NMR. High‐resolution 1H and 13C‐NMR spectra were used to study the microstructure of the copolymer. As an example, the three well‐resolved carbonyl resonances in the 13C‐NMR spectrum were assigned to the MATRIF‐centered triads VMV, VMM, and MMM, respectively, (V and M stand for VCN and MATRIF, respectively). The presence of VCN dyads (e.g., in VVM and VVV sequences) was shown to be marginal or absent altogether. Thermogravimetric analysis of poly(VCN‐co‐MATRIF) copolymer showed good thermal stability, and its main pyrolytic degradation taking place only above 368 °C. A 4% weight loss at about 222 °C suggested the presence of a few VCN homodyads, possibly inducing thermal depolymerization. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

15.
2-Trimethylsilyloxy-1,3-butadiene (TMSBD), the silyl enol ether of methyl vinyl ketone, was homopolymerized with a radical initiator to afford polymers with a molecular weight of ca. 104. Radical copolymerizations of TMSBD with styrene (ST) and acrylonitrile (AN) in bulk or dioxane at 60°C gave the following monomer reactivity ratios: r1 = 0.64 and r2 = 1.20 for the ST (M1)–TMSBD (M2) system and r1 = 0.036 and r2 = 0.065 for the AN (M1)–TMSBD (M2) system. The Q and e values of TMSBD determined from the reactivity ratios for the former copolymerization system were 2.34 and ?1.31, respectively. The resulting polymer and copolymers were readily desilylated with hydrochloric acid or tetrabutylammonium fluoride as catalyst to yield analogous polymers having carbonyl groups in the polymer chains.  相似文献   

16.
The influence of the total monomers concentration and of the copolymerization solvent on the reactivity ratio, r1, of methyl methacrylate (MMA) (M1) ω-(p-vinylbenzyl ether) macromonomer of poly(2,6-dimethyl-1,4-phenylene oxide) (PPO–VBE) (M2) monomer pair was investigated. For two different molecular weights of PPO-VBE macromonomers: M n = 15,200 and M n = 5,100, the determined reactivity ratio, r1, decreases with the increase in macromonomer concentration. Therefore the reactivity of the macromonomer, 1/r1, follows the opposite trend. The influence of monomers concentration on r1 is higher for higher molecular weight macromonomers. The nature of polymerization solvent also affects the value of reactivity ratio, r1. Micelle formation was demonstrated by 1H-NMR spectroscopy performed on the resulting graft copolymers in different solvent mixtures. An attempt to explain the observed concentration and solvent effects based on the partition of comonomer concentrations between the bulk of solvent and around the growing chain is presented. Based on this explanation, the determined r1 represents a product of the partition coefficient, k, and the true reactivity ratio, r10.  相似文献   

17.
Abstract

Radical homopolymerization of N-[4-N′-(α-methylbenzyl)-aminocarbonylphenyl]maleimide ((S)-MBCP) was carried out at 50 and 70°C for 24 h to give optically active polymers ([α]25 D = 159.8 to 163.4°). Radical copolymerizations of (S)-MBCP (M1) were performed with styrene (ST, M2, methyl methacrylate (MMA, M2) in THF at 50°C. The monomer reactivity ratios (r 1, r 2) and the Alfrey-Price Q, e values were determined as follows: r 1 = 0.32, r 2= 0.14, Q 1 = 1.74, e 1 = 0.96 in the (S)-MBCP-ST system; r 1 = 0.54, r 2 = 0.93, Q 1 = 1.11, e 1 = 1.23 in the (S)-MBCP-MMA system. Chiroptical properties of the polymers and the copolymers were also investigated, and asymmetric induction into the copolymer main chain is discussed.  相似文献   

18.
The spontaneous copolymerization of N-phenylmaleimide (NPMI) (M1) with ethyl α-phenylacrylate (EPA)(M2) were carried out in dioxane at 85°C. A high alternating tendency was observed. The monomer reactivity ratios were r1 = 0.07 ±0.01 and r2 = 0.09 ± 0.02. The maximum copolymerization rate and molecular weight occurs at 70–80 mol% (M1) in feed ratio. The spontaneous alternating copolymerization is considered to be carried out via a contact-type charge transfer complex (CTC) formed between the monomers. Thermogravimetric analyses (TGA) indicate the resulting copolymers have high thermal stability. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 2927–2931, 1998  相似文献   

19.
The copolymers prepared in this study by free radical copolymerization of N-vinylpyrrolidone (M 2) with 4-vinylbenzylchloride (M 1) using 2,2′-azobisisobutyronotrile (AIBN) initiator in 1,4-dioxane solvent at 70°C were characterized by FTIR, 1H-NMR and 13C-NMR techniques. Polymer solubility was tested in both polar and nonpolar solvents. The thermal properties were studied by thermogravimetric analysis (TGA) and differential scanning calorimeter (DSC). Copolymer compositions were established by H1-NMR spectra, while reactivity ratios of the monomers were computed using the linearization methods viz., Fineman-Ross (FR) (r 1 = 1.67 and r 2 = 0.67), Kelen-Tudos (KT) (r 1 = 1.77 and r 2 = 0.65) and extended Kelen-Tudos (EK-T) (r 1 = 1.72 and r 2 = 0.63) methods at lower conversion. Furthermore, reactivity ratios in nonlinear error-in-variables method (RREVM) also compute the reactivity ratios (r 1 = 1.76 and r 2 = 0.66); these are found to be in good agreement with each other. The distribution of monomer sequence along the copolymer chain was calculated using a statistical method based on the calculated reactivity ratios.  相似文献   

20.

A new methacrylic monomer, 4‐nitro‐3‐methylphenyl methacrylate (NMPM) was prepared by reacting 4‐nitro‐3‐methyl phenol dissolved in methyl ethyl ketone (MEK) in the presence of triethylamine as a catalyst. Copolymerization of NMPM with methyl methacrylate (MMA) has been carried out in methyl ethyl ketone (MEK) by free radical solution polymerization at 70±1°C utilizing benzoyl peroxide (BPO) as initiator. Poly (NMPM‐co‐MMA) copolymers were characterized by FT‐IR, 1H‐NMR and 13C‐NMR spectroscopy. The molecular weights (Mw and Mn) and polydispersity indices (Mw/Mn) of the polymers were determined using a gel permeation chromatograph. The glass transition temperatures (Tg) of the copolymers were determined by a differential scanning calorimeter, showing that Tg increases with MMA content in the copolymer. Thermogravimetric analysis of the polymers, performed under nitrogen, shows that the stability of the copolymer increases with an increase in NMPM content. The solubility of the polymers was tested in various polar and non‐polar solvents. Copolymer compositions were determined by 1H‐NMR spectroscopy by comparing the integral peak heights of well separated aromatic and aliphatic proton peaks. The monomer reactivity ratios were determined by the Fineman‐Ross (r1 =7.090:r2=0.854), Kelen‐Tudos (r1=7.693: r2=0.852) and extended Kelen‐Tudos methods (r1=7.550: r2= 0.856).  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