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 93 毫秒
1.
The chemical homogeneity of a series of copolymers obtained by nucleophilic substitution of organolithium reagents RCH2Li [R = S(C6H5), SOCH3, SO2CH3 and SO2N(CH3)2] on a high molecular weight poly(methyl methacrylate) (PMMA, DPn = 700) has been studied by different methods, over a wide range of substitution degrees (0.14 ? DSm ? 0.76). ‘Cross’ fractionation is much more efficient than ‘one direction’ fractionation, and it allows the determination of \?gs2 variance values as low as 2 × 10?4 with sufficient accuracy. The light scattering method is far less sensitive in this range of low compositional polydispersity, even for a suitable system. The fairly high chemical homogeneity observed for all the copolymers, prepared either in homogeneous or heterogeneous solution, may be correlated with both the high molecular weight of the PMMA precursor and the autoretarded kinetics of the substitution process.  相似文献   

2.
Thirteen fractions of poly(phenyl acrylate) have been prepared with weight-average molecular weight ranging from 0.158 × 106 to 2.57 × 106 g mol?1. The temperature coefficient of the unperturbed dimensions and the glass transition temperature were found to be ?1.8 × 10?3 deg?1 and 55.6°C respectively. Good accord was obtained among different methods for establishing θ-conditions of 11.5°C in ethyl lactate. From viscometry, osmometry and light scattering under θ-conditions, as well as in a good solvent, the unperturbed dimensions were determined via several procedures yielding a value of [〈r20wM?w]12 = 6.0 (±0.2) × 10?9cm g?12mol12. This corresponds to a steric factor υ = 2.37 (±0.08) and a characteristic ratio C = 11.3 (±0.8). The polymer chain is thus more rigid than poly(methyl acrylate), but less rigid than poly(phenyl methacrylate). With respect to its Tg and flexibility, poly(phenyl acrylate) bears a strong similarity to poly(benzyl methacrylate).  相似文献   

3.
Walther Burchard 《Polymer》1979,20(5):589-592
Relationships are given between the z-average radii moments r?nz and the common moments r?n of a size distribution. Instructions are given for finding the type and width of a size distribution from measurements of the r?nz moments.  相似文献   

4.
The synthesis and characterization of methacrylate-ended macromers (M?n 500 to 10 000) and their copolymerization with styrene (M2) is described. The experimental errors in the values of the reactivity ratios r1 render them meaningless. Values of r2 can be determined with more precision and increase from 1.06 to 1.55 as the molecular weight of the macromer increases. This behaviour is due to steric effects, not diffusion-controlled propagation. It is shown that the assumptions that 1 > r1[M1][M2] and r2 >[M1][M2] are only valid for macromers of M?n > ca. 10 000.  相似文献   

5.
K. Dodgson  D. Sympson  J.A. Semlyen 《Polymer》1978,19(11):1285-1289
A preparative gel permeation chromatographic (g.p.c.) instrument has been constructed and used to separate broad fractions of cyclic poly(dimethyl siloxanes) into sharp fractions with heterogeneity indices M?wM?n = 1.05 ± 0.02. The number-average molecular weights M?n of the cyclic polymer fractions obtained were as high as 50 000, corresponding to number-average numbers of skeletal bonds n?n up to 1300. The concentrations of linear poly(dimethyl siloxanes) in all but the highest molecular weight cyclic polymer fractions prepared are believed to be negligible. The preparative g.p.c. instrument was also used to obtain some sharp fractions of linear poly(dimethyl siloxanes).  相似文献   

6.
Di- and tri-block copolymers of ethylene oxide and styrene function as effective phase transfer catalysts in the reaction of solid potassium phenoxide with n-butylbromide in refluxing toluene to give virtually quantitative yields of n-butyl phenyl ether. A wide range of copolymer structures have been examined and the catalytic activity is found to increase with M?n of both the styrene and the ethylene oxide segments levelling off at M?n styrene block ~30 000 and M?n ethylene oxide block ~60 000. Beyond M?n ethylene oxide block ~100 000 rates of reaction drop again towards the value for a high molecular weight homopolymer of ethylene oxide. Kinetic analysis suggests the rate controlling process to be the bimolecular reaction between complexed potassium phenoxide and n-butyl bromide, and the activation energy for the reaction is the same as that for reactions catalysed by low molecular weight oligoethers. From the kinetic dependence of the concentration of copolymer catalysts, the dependence on the structure of individual copolymers and from the known physical behaviour of these copolymers in toluene solution, catalysis appears to involve micellar aggregates of copolymer chains.  相似文献   

7.
Poly(ethyl acrylate) (PEA), solution polymerized in methyl ethyl ketone by free radical initiation, was fractionated and the fractions were characterized by light scattering, viscometry and osmometry. Fractions obtained were in the molecular weight range of 0·3 × 106 to 1·6 × 106 with a polydispersity of 1.40. The following Mark-Houwink relations were established:
[η]35°Cacetone =4·15×10?2M0?61W
[η]35°CMEK =2·03×10?2M0?66W
[η]39.5°Cn-propanal =7·89×10?2M0?50W
It was found that n-propanol at 39.5°C was a theta solvent for poly(ethyl acrylate) and that acetone was a poor solvent compared to methyl ethyl ketone. A relation between the molecular dimension and the molecular weight was established. It was observed that the chain dimensions of poly(ethyl acrylate) and poly(butyl acrylate) were considerably larger than poly(ethyl methacrylate) and poly(butyl methacrylate) respectively. The validity of various extrapolation procedures that have been proposed for calculating the unperturbed dimensions have been examined. The steric factor for PEA was 2·16 compared to 2·10 for poly(ethyl methacrylate). Root mean square end-to-end distances were calculated from the Debye-Bueche and Kirkwood-Riseman methods and compared with the experimental values.  相似文献   

8.
Dilute solution properties of poly(phenyl thiolmethacrylate) (PTPh) and poly(o-methylphenyl-thiolmethacrylate) (PTMPh) were studied by gel permeation chromatography, light scattering, osmotic pressure and viscosity measurements in different solvents. Relations between intrinsic viscosity [η], z-average root-mean-square end-to-end distance (r?2)12 and molecular weight M?w were established. The unperturbed dimensions were calculated by different methods using M?w and [η]. PTPh (σ = 2.26) and PTMPh (σ = 2.27) have practically the same flexibility, but both are more flexible than poly(phenyl methacrylate) (σ = 2.50). In our opinion, this is due to the change in the local intramolecular interaction and the higher flexibility of the sulphur-containing side group.  相似文献   

9.
Several polystyrene-poly(ethylene oxide)-polystyrene (PS-PEO-PS) and poly(methyl methacrylate)-polystyrene-poly(methyl methacrylate) (PMMA-PS-PMMA) block copolymers, synthesized by free-radical polymerization, were studied in various solvents by using a light-scattering technique. The copolymers, which have different lengths of central blocks, had molecular weights within the range 3.0 × 104 to 1.6 × 106. It was found that almost all of them were confirmed as block copolymers from the variation of the product Mapp〈S2app12 with WAvAv, although they were rather heterogeneous. The copolymer compositions determined either from the additivity of the refractive index increments of its constituent parts and the copolymer or from ultra-violet analysis were in excellent agreement with each other.  相似文献   

10.
The dimensions of both cyclic and linear poly(dimethyl siloxanes) in dilute solution in benzene-d6 have been measured by small-angle neutron scattering. The mean-square radii of gyration of the linear polymers are consistent with values predicted from published data, including experimental molar cyclization equilibrium constants. The average dimensions of the cyclic poly(dimethyl siloxanes) in fractions containing z-average numbers of bonds n?z in the range 130 < n?z < 550, were found to be considerably smaller than those of the corresponding linear polymers. The neutron scattering results give a value for the ratio of the z-average radii of gyration for linear and ring poly(dimethyl siloxanes) (containing the same number of monomer units) 〈s2z,l<s2z,r = 1.9 ± 0.2. This ratio may be compared with the value of 2.0 predicted theoretically for ‘flexible’ high molecular weight linear and cyclic polymers, unperturbed by excluded volume effects.  相似文献   

11.
Ultramicroscopy studies have been made of micelle formation by two poly(styrene)-poly(isoprene) block copolymers in organic solvents (N,N-dimethylacetamide and n-decane respectively) and a poly(l-glutamic acid)-poly(l-leucine) block copolymer in an aqueous solution of 0.2M NaCl at pH = 4.0. The technique provides a method of determining the number-average translational diffusion coefficient) D?n, of association colloids and leads, via the Stokes-Einstein relation, to a measure of the number-average of the reciprocal hydrodynamic radius (RD?1)n for spherical particles. Particles having a radius less than approximately 30 nm were too small to be detected by the technique.The ultramicroscopy results were compared with data obtained by laser light scattering photon correlation spectroscopy which provides a measure of the z-average translational diffusion coefficient. D?z. An additional comparison was made by carrying out measurements on two well-characterized poly(styrene/divinyl benzene) latices.  相似文献   

12.
Yasuhiko Onishi 《Polymer》1980,21(7):819-824
Effects of the molecular weight of dextran on its graft copolymerization with methyl methacrylate (MMA), initiated by ceric ammonium nitrate (CAN), have been investigated. The results indicate that grafting (%), graft polymerization (%) (ψ), the overall rate constant (k′) for consumption of Ce4+, and branch PMMA were influenced significantly by the molecular weight of the backbone polymer dextran. The number of branch PMMA chains per dextran molecule was 0.05 ~ 0.30 for M?w 9000 dextran (D1), 0.35 ~ 0.55 for M?w 61 000 (D2), and 0.8 ~ 1.6 for M?w 196 000 (D3), respectively. The relationship between the rate of graft polymerization and M?w (the weight-average molecular weight of dextran) was expressed by the equation: Rpg = ?AlogM?w + B. Another linear relationship was obtained between In (100 ? ψ) and reaction time (t) for both D1 and D2 samples or In t for D3. Detailed kinetic analysis has been made on the basis of the latter relationship. Mechanical properties were also studied on the moulded sample plates of these copolymers.  相似文献   

13.
Polymerization, and copolymerization with styrene, of m,p-chloromethylstyrene have been carried out at 75°C, in chlorobenzene and in the presence of AIBN ([AIBN] ? 6 × 10?2, and 12 × 10?2m, respectively). The polymer molecular weights, determined by g.p.c., are: M?w = 8670, M?n = 5860, and M?w/-Mn = 1.48 for the homopolymer, poly(m,p-chloromethylstyrene), (1a); and M?w = 8805, M?n = 5144, and M?w/-Mn = 1.71 for the copolymer, copoly(m,p-chloromethylstyrene-styrene), (2a). A series of phosphine derivatives of both 1a and 2a are prepared by the reaction of the polymers with either chlorodiphenylphosphine/lithium, or diphenylphosphine/potassium tert. butoxide. A number of other potentially electroreactive derivatives of 2a are obtained by reacting the polymer with 2-aminoanthraquinone, 3-N-methylamino-propionitrile, or 2-(2-aminoethyl) pyridine. The phosphinated polymers are reacted with bis-benzonitrilepalladium-(II) chloride to obtain a series of polymer-palladium(II) complexes containing 8.5–12.9% palladium. Similarly, reaction of the last-named bidentate polymeric ligand with cupric acetylacetonate, or cupric sulphate pentahydrate, produces polymer-copper(II) complexes having 5.8, or 3.3% copper, respectively. The inter/intra-chain nature of some of the side reactions during the derivatization of the chloromethylated polymers, and that of the complex formation between transition metal centres and macromolecular ligands, are briefly discussed in view of the experimental results.  相似文献   

14.
The isothermal crystallization of poly(ethylene-terephthalate) (PETP) fractions, from the melt, was investigated using differential scanning calorimetry (d.s.c.). The molecular weight range of the fractions was from 5300–11750. Crystallization temperatures were from 498–513 K. The dependence of molecular weight and undercooling on several crystallization parameters has been observed. Either maxima or minima appear at a molecular weight of about 9000, depending on the crystallization temperature. The activation energy values point to the possibility of different mechanisms of crystallization according to the chain length. A folded chain process for the higher M?n chains and an extended chain mechanism for the lower M?n chains. The values of the Avrami equation exponent n vary from 2–4 depending on the crystallization temperature; non-integer values are indicative of heterogeneous nucleation. The rate constant K depends on Tc and M?n, showing maxima related to the Tc used. The plot of log K either vs. (ΔT)?1 and (ΔT)?2 or TmT(ΔT) and T2mT(ΔT)2 is linear in every case.  相似文献   

15.
S. Franco  A. Leoni 《Polymer》1973,14(1):2-4
The influence of different solvents on the copolymerization behaviour of N-(1,1-dimethyl-3-oxobutyl)-acrylamide (diacetone acrylamide, DAAM) with styrene and methyl methacrylate has been investigated. The calculated relative reactivities of DAAM towards the polystyryl and poly(methyl methacrylyl) radicals (1r2) are practically unaffected by the reaction medium; on the contrary, the relative reactivities of both styrene and methyl methacrylate monomers towards the poly(diacetone acrylamide) radical (1r1) are influenced by the reaction solvents. This effect, however, is less significant than that found previously for N-methyl acrylamide copolymerization.  相似文献   

16.
M. Kučera  Z. Salajka  P. Hudec 《Polymer》1985,26(11):1733-1736
The efficiency as well as the feasibility of a transformation of ionic polymerization centres to those possessing an opposite charge, has been examined. The transformation was effected via the interaction of the polymeric dianion (dication) being transformed, with a molar excess of an oligomeric dication (dianion). The result of the interaction was evaluated by measuring the molecular weight of the product. The transformation of anionic ends to cationic ends (or the reverse) proceeds smoothly at a twofold excess of the transforming diion. When this excess is lower then combination of oppositely charged ends of both reactants takes place simultaneously with transformation. The transformation of dianionic living polystyrene (Mn=5000) (dicationic poly(tetramethylene oxide), M?n=6000) to dicationic (dianionic) product deviates only slightly from the theoretically assumed course; the transformation of diions of higher molecular weight (M?n=17–30×103) to polymers with oppositely charged ends obeys reasonably the assumed kinetics only at a two-fold excess of the transforming oligomer; below this ratio differences are more pronounced. The polymerization of suitable monomers on the transformed centres gives rise to block copolymers. Detailed structures of these copolymers have not yet been elucidated due mainly to difficulties in separation of block and homopolymers from mixtures.  相似文献   

17.
K. Dodgson  J.A. Semlyen 《Polymer》1977,18(12):1265-1268
The limiting viscosity numbers of ten cyclic and ten linear poly(dimethyl siloxane) fractions have been measured in a π-solvent (butanone at 293K) and in two ‘good’ solvents (toluene and cyclohexane at 298K). The dimethyl siloxane fractions studied were in the molecular weight range 800 < M?w < 17 000. The data obtained are compared with related studies published in the literature. The ratio of the limiting viscosity numbers [η]r and [η]l of the cyclic and linear poly(dimethyl siloxanes) with M?w > 2500 was found to be 0.67 in butanone at 293K. This value is identical (within experimental error) to the theoretical ratio [η]r[η]l = 0.66 predicted by Bloomfield and Zimm and others for ring and chain polymers in π-solvents. The ratio [η]r[η]l was found to be somewhat smaller for the higher molecular weight polymers in the ‘good’ solvents.  相似文献   

18.
E Ikada  T Sugimura  T Aoyama  T Watanabe 《Polymer》1975,16(2):101-104
Dielectric properties of vinyl acetate and methyl methacrylate oligomers were studied in order to compare the dielectric properties of an oligomer with those of the corresponding high polymers. The two oligomers showed asymmetric dielectric relaxations at room temperatures. The complex dielectric constants vs. angular frequencies for these oligomers were well represented by the Havriliak-Negami equation:
?ast;??=?0??[1+(jωτ0)1?α]β
The distribution parameters (1 - α) and β of vinyl acetate and methyl methacrylate oligomer are almost equal to those of poly(vinyl acetate) and poly(methyl methacrylate), respectively. It was concluded that the distribution of relaxation times was independent of the molecular weight of the polymers.  相似文献   

19.
Attempts were made to correlate the average size of the network of crosslinked vinyl acetate-glycidyl methacrylate (GMA) copolymer to the maximum size of permeable molecules when the copolymers were used as g.p.c. packing materials. The root-mean-square of the end-to-end distance (r?2)12 of the hydrodynamic radius (s?2)12, which characterizes the maximum extent of the molecules, were calculated. Meanwhile, the average size of the copolymer network, rc, was estimated from the average molar weight between crosslinkages, this being calculated from the equilibrium rubber elastic modulus, Er, obtained by dynamic discoelastic measurements. It was found that the maximum size which can permeate the copolymer decreases as the content of GMA in the copolymer increases. From viscoelastic measurements, Er was found to increase as the content of GMA in the copolymer increases. Further, linear relationships were obtained among rc, (r?2)12, and (s?2)12 covering the wide range of the copolymer composition.  相似文献   

20.
The effects of temperature and catalyst homogeneity on the molecular weight distribution (MWD) and stereochemical regulation of polypropylenes produced by Ti(OC4H9)4Al2(C2H5)3Cl3 system have been investigated. The MWD of polymers obtained at temperatures below 21°C were unimodal and narrow (M?wM?n?2.0), whereas those obtained at temperatures higher than 31°C were bimodal with one narrow distribution and the other broad one (M?wM?n=18) at higher molecular weights. The existence of two different types of catalyst, one soluble with homogeneous catalytic centres and the other insoluble with heterogeneous catalytic centres was found in the polymerization at 41°C. At temperatures below 21°C only soluble catalyst was present and produced isotactic polypropylenes with [m]=0.65. The isospecific nature of soluble titanium-based catalyst is greatly contrasted to the syndiospecific nature of soluble vanadium-based catalyst.  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