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 78 毫秒
1.
The statement is often made in the polymer literature, without proof, that M zM wM n, where M z, M w, and M n are the z-, z weight-, and number-average molecular weights respectively. Four proofs of a generalization of these inequalities are given. It is shown that a higher-order molecular weight average is larger than a lower-order one, regardless of the form of the molecular weight distributions, except for the case when all the molecules have the same molecular weight. A brief discussion of the viscosity-average molecular weight is also included.  相似文献   

2.
Poly(α-methylstyrene-b-isobutyl vinyl ether-b-α-methylstyrene) triblock polymers have been prepared by blocking α-methyl-styrene (αMeSt) from biheaded quasiliving poly(isobuty1 vinyl ether) (PIBVE) cations generated with the bifunctional p-dicumyl chloride/AgSbF6 initiating system in methylene chloride solvent at -90°C. The products were fractionated with 2-propanol, a good solvent for PIBVE and a nonsolvent for PaMeSt. The 2-propanol-insoluble fractions had much higher molecular weights (M n = 30,500–69,100) than the starting PIBVE (M n =6,600–10,600) and contained 13–29 wt% IBVE together with 87–71 wt% αMeSt units. The 2-propanol-soluble fractions (M n = 7,300–11,600) contained ~90 wt% IBVE and ~10 wt% αMeSt units.  相似文献   

3.
The molecular weight distribution (MWD) of crystallizable polyphenylacetylene prepared near room temperature in the presence of ferric acetylacetonate and triethylaluminum was determined through use of fractions characterized by vapor pressure osmometry and gel permeation chromatography (GPC). The number- and weight-average molecular weights (M n and M w) are both less than the molecular weight corresponding to the maximum of the weight distribution function, which lacks a high molecular weight tail. M wandM n is less than is consistent with models allowing for chain termination characteristic of vinyl polymers. GPC elution volumes are much less than those characteristic of polystyrene of similar molecular weight, and the Mark-Houwink exponent is high (2.4 for M v 4800 to 6800). These data indicate more rodlike behavior than for polystyrene of equivalent molecular weight. The MWD and other data suggest intramolecular chain termination, possibly associated with the molecule's tendency to form paramagnetic defect states.  相似文献   

4.
Quasiliving carbocationic polymerization of methyl vinyl ether (MVE) was achieved with the p-dicumyl chloride (p-DCC)/AgSbF6 initiator system by the slow and continuous monomer-addition (quasiliving) technique. A polar solvent (CH2Cl2) and a low reaction temperature (-70°C) were optimum for the quasiliving MVE polymerization. Under these conditions, the number-average molecular weight (M n) of poly(MVE) increased linearly with the cumulative weight of added monomer (WMVE), and linear M n versus WMVE plots passed through the origin. M n's were inversely proportional to the initial initiator (p-DCC) concentration. Reactions in a nonpolar solvent (toluene) at -70°C or in a polar solvent (CH2Cl2) at ?30°C resulted in deviations from these quasiliving characteristics. Block polymerization of MVE from quasiliving poly(isobutyl vinyl ether) dications by the quasiliving technique (p-DCC/AgSbF6 initiator, CH2Cl2 solvent,(-70°C) led to novel isobutyl vinyl ether (IBVE)-MVE block polymers in high yield (>93 wt%) and at high blocking efficiency. The block polymers, most likely poly(MVE-b-IBVE-b-MVE), having M n = 10,900–14,000 [M n(center block) = 6,200–9,0001, were soluble in n-heptane and insoluble in water, and gave hazy homogeneous solutions when dissolved in methanol at room temperature.  相似文献   

5.
Graft copolymers of poly(diallyldimethylammonium chloride), (poly-DADMAC), with acrylamide were synthesized using a ceric salt/nitric acid initiation system. The effects of concentration of initiator, monomer, and substrates were studied. Copolymers were characterized by viscometry and size-exclusion chromatography. The highest molecular weight ( M w ) of graft copolymer obtained was 1.70 × 106. The compositions of copolymers are dependent upon the ratios of the concentration of monomer to the concentration of substrate. The highest content of DADMAC monomer unit in the copolymer was 33 wt%.  相似文献   

6.
A detailed analysis of elementary reactions of carbocationic polymerization culminated in the prediction and subsequent experimental demonstration of quasiliving polymerization. Quasiliving polymers are formed in a system provided that the process of chain termination and chain transfer to monomer are absent or reversible, i.e., the propagating ability of the chain end is maintained throughout the experiment, and the molecular weight increases in proportion to the cumulative amount of monomer added. The chain end can be active (carbocation) or dormant (reactivable polymeric olefin or cation source). Chain transfer is suppressed by keeping the monomer concentration low. Quasiliving polymerizations are maintained by continuous slow feeding of dilute monomer to a charge containing the initiating or propagating species (quasiliving polymerization technique). A comprehensive kinetic scheme has been developed that describes quasiliving polymerization in quantitative terms. Quasiliving polymerization was demonstrated experimentally in the “H2O”/BCl3/α-methylstyrene and cumyl chloride/BCl3/α-methylstyrene systems. M n versus monomer input plots are linear over wide ranges, indicating quasiliving conditions, and poly(α-methylstyrenes) with M n > 2 × 105 have been obtained, Molecular weight distributions were found progressively to narrow and dispersion ratios M w/M n to decrease.  相似文献   

7.
Poly(meta-aryl sulfide amide amide) (m-PASAA) was prepared with aromatic nucleophilic substitution reaction: by the step polycondensation of sodium sulfide(Na2S· xH2O) with 3,3′ -bis(4-diflurobenzoyl) aryl diamine between 180–202°C at atmospheric pressure. The polymers were characterized by FT-IR spectrum, 1H-NMR spectrum, 13C-NMR spectrum, X-ray diffraction, element analyzer, DSC, TGA, AFM, instron universal tester and dissolvability experiment. The intrinsic viscosity of m-PASAA was 0.41–0.46 dl/g obtained with optimum synthesis conditions. The polymers were found to have excellent thermal performance with glass transition temperature (Tg) of 233.5–277.8°C, initial degradation temperature (Td) of 447–456.7°C. They could afford flexible and strong films with tensile strengths 38.4–46.1MPa. At the same time, their solubility was much better than polyphenylene sulfide (polyphenylene sulfide scarcely dissolves in whole organic solvents under 200°C (1 Yang, J. 2006. PAS resin and its application, China: Chemical Industry Press.  [Google Scholar])).  相似文献   

8.
Abstract

A two-stage process was developed for the living polymerization of isobutylene (IB) employing di-tert-alcohol initiators in conjunction with BCl3 coinitiator in the first or initiation stage, followed by TiCl4 coinitiator in the second or propagation stage; the process was shown to yield high molecular weight (up to M n 20,000), narrow molecular weight distribution (MWD) M w/M n = 1.1–1.2) di-tert-chlorine telechelic polyisobutylenes (tCl-PIB-Clt). The initiation stage involves the homogeneous solution living polymerization of IB induced by the di-tert-alcohol/BCl3 combination in the presence of an electron donor such as N,N-dimethylacetamide in CH3Cl solvent at ?80°C and proceeds up to M n < 5000; this is followed by the propagation stage in which TiCl4 and the bulk of IB plus a sufficient amount of n-C6H14 are added to the charge to bring the solvent composition to CH3Cl/n-C6H14 60/40 v/v and the living polymerization is continued until high M n product is obtained. This two-stage process was developed because 1) it employs very inexpensive chemicals; 2) di-tert-alcohol/BCl3 combinations initiate living IB polymerization in CH3Cl but the product after reaching M n ≤ 5000 precipitates out of the CH3Cl solution, and di-tert-alcohol/BCl4 combinations do not initiate IB polymerization; and 3) di-tert-alcohol/BCl3 systems do not initiate (or only very slowly) the living polymerization of IB in CH3Cl/n-C6H14 mixtures, whereas similar TiCl4-based systems do. The polymerization remains living during both stages although the propagating species and solvent polarity are profoundly altered. The livingness of the system has been analyzed by kinetic experiments and the structure of the tCl-PIB-Clt product by routine spectroscopic means.  相似文献   

9.
A new method for the determination of the sedimentation coefficient-molecular weight relation is proposed. Based on the very low sensitivity of the corresponding sedimentation coefficient average S 0,1 and the weight-average molecular weight M w (calculated according to the generalized Flory-Mandelkern equation) to changes of the a parameter, the S0,1 and M w values are estimated from constant initial values of the as parameter for all those polymer-solvent systems for which the real as values do not exceed the interval 0.3-0.5. The approximations involved give an error lower than 8.5%, Le., below the experimental errors of the Mw values determined for polydiSDerse samDles. The new method of determining the So -M relation was applied to the system styrene-acrylonitrile copolymer (22.6-wt%oacrylonitrile content) in acetone at 25°C and yielded the following relation: S0=2.90 × 10 ?15 M 0,49 sec. Although in the case of this polymer-solvent system the a value of 0.49 is close to the one corresponding to θ systems, the method is shown by model calculations to be of general applicability and especially useful in the case of nonideal polymer-solvent systems.  相似文献   

10.
A simple reusable apparatus for the synthesis of up to 40 g quantities of poly(styrene-b-isoprene) diblock copolymers of reasonably low (1.2 to 1.5) polydispersity has been described. The diblock copolymers synthesized were characterized by gel permeation chromatography (GPC), membrane osmometry, viscosimetry, and nuclear magnetic resonance (NMR) spectroscopy. Number-average molecular weights (M n) calculated from the raw GPC chromatographs of the diblock copolymers using the summation method and M versus elution volume plots for polystyrene and polyisoprene standards agree well with those measured experimentally with osmometry. It is suggested that for polydisperse block copolymers this method is simpler than the use of a universal calibration curve. Mark-Houwink constants K ans a for polyisoprene having 18% (1,2-), 66% (3,4-), and 16% (1,4-) microstructure were found to be 3.2 × 10?4 dL/g and 0.67, respectively, in THF at 25°C. In toluene at 30°C, K = 2.0 × 10?4 dL/g and α = 0.7 were obtained. The diblock copolymers had 26% (1,2-), 60% (3,4-), and 14% (1,4-) microstructure in the isoprene segments, and the values of K and a for these copolymers (PS > 50%, M 20.0 × 103) in THF at 25°C were 9.0 × 10?5 dL/g and 0.75. For M < 20.0 × 103 the value of α was 0.5. The experimental values of [η] were found to be lower than those calculated theoretically, presumably due to the polydisperse nature and the biellipsoidal configuration of the diblock copolymers.  相似文献   

11.
Abstract

Estimation of molecular weights from GPC data is complicated when the polymer sample consists of a mixture of homopolymers or of statistical copolymers with nonuniform compositions. This is because sizes of solvated polymer coils depend on solvent interaction with both the homo-and hetero-units of the copolymers and because the extent of solvation of different homopolymers can differ. The overall degree of solvation may change effectively with composition and use of a single “average” set of Mark-Houwink constants in calibration procedures will then produce false molecular weight data from the GPC data. A new molecular weight average, M x, is defined to overcome this problem. This average can be determined from the GPC chromatogram and intrinsic viscosity of the sample in the GPC solvent. Mark-Houwink coefficients are not needed. M x lies between M w and M z.  相似文献   

12.
Abstract

Copolymers of p-nitrobenzyl acrylate and methyl acrylate with different feed ratios are synthesized in ethyl methyl ketone using benzoyl peroxide as a free radical initiator at 70 ± 1°C. The copolymers were characterized by IR and 1H-NMR spectroscopic techniques. Copolymer compositions were determined by 1H-NMR analysis of the polymers. The monomer reactivity ratios were determined by the application of conventional linearization methods such as Fineman–Ross and Kelen–Tüdös. Gel permeation chromatography was used for determining the molecular weights M n and M w, and the polydispersity index. The intrinsic viscosities and the thermal properties of the homo- and copolymers are also discussed.  相似文献   

13.
Synthesis of aromatic poly(ether ketone) (3) with a narrow molecular weight distribution (Mw/Mn) was investigated via polycondensation. Mns of 3 could be controlled varying the feed ratio of monomer (1) and initiator (2) maintaining relatively narrow Mw/Mns (<1.3). The kinetics of polycondensation obeyed a first-order relationship between polycondensation time and -(1/[2]0) ln([1]/[1]0), and the rate of polycondensation was estimated as 2.57 mol−1 L h−1. MALDI-TOF mass analysis of 3 indicated that polycondensation should proceed via chain growth manner to give 3 having an initiator unit in each chain end.  相似文献   

14.
Abstract

Ultrasonic (70 W, 20 kHz) solution (2% THF) degradations of polystyrene (PS), poly(α-methylstyrene) (PαMeS), poly(p-isopropyl α-methylstyrene) (PpiPrαMeS), poly(p-chlorostyrene) (PpCIS), poly(p-bromostyrene) (PpBrS), and poly(p-methoxystyrene) (PpOMeS) have been carried out in tetrahydrofuran at 27° C. The average number of chain scissions S (where S = [(M n)0/(M n)t] - 1), computed from the overall values of [(M n)0 and (M n)t, were found to be different from those of S' (where S' = α([(M n)0/(M n)t] - 1)) based on the component (only that part of the polymer which is involved in degradation) data of the weight fraction (α), (M n)0, and (M n)t), S' for polystyrene and substituted polystyrene follows the order PS gt; PpCIS gt; PpiPrαMeS gt; PpBrS gt; PpOMeS gt; PαMeS. In the case of PS where degradations were also carried out at -20°C, lowering of the temperature increased the weight fraction of polymer degraded as well as S. Based on the viscosity and GPC data, it is concluded that the ultrasonic solution degradation of PS does not lead to branched polymers.  相似文献   

15.
Abstract

Living copolymerization of the isobutylene (IB)-p-methylstyrene (pMeSt) monomer pair in combination with the constant copolymer composition (CCC) technique produces high molecular weight ( M n ≈ 100,000 g·mol?1) and narrow molecular weight distribution ( M w/ M n ≈ 1.45) compositionally uniform IB/pMeSt copolymer molecules in the industrially important IB/pMeSt = 97–99/3–1 mol% composition range. Syntheses were carried out with TiCl4 coinitiator in n-butyl chloride homogeneous solution at ?85°C by the use of the Leidenfrost reactor (i.e., by direct cooling of the charge with liquid nitrogen). In order to carry out the CCC technique it was necessary to obtain reliable copolymerization reactivity ratios. These investigations led to rIB = 0.5 ± 0.1 and r pMeSt = 10 ± 4. The attainment of CCC and living copolymerization conditions has been quantitatively demonstrated by dedicated diagnostic plots. Specifically, the attainment of CCC conditions was proven by the analysis of composite rate plots (comonomers input and corresponding copolymer formed versus time) and composition plots (comonomer composition in feed and copolymer formed versus weight of copolymer formed, W p), and living copolymerization was proven by linearly ascending number-average molecular weight of copolymer ( M n) versus W p plots starting at the origin.  相似文献   

16.
17.
Abstract

Pectic polysaccharides (HP, RP) were extracted from Mongolia hawthorn berries and from rhubarb stalks by hot aqueous oxalic acid treatment. Both polysaccharides contained approximately 78% of galacturonate. HP displayed a low degree of esterification (DE) and arabinose/galactose molar ratio 1.7:1. RP was higher in DE and contained these sugars in a molar ratio 0.5:1. Rhamnose, fucose, xylose, glucose, mannose, 2-O-methylxylose, and 2-O-methylfucose were found in lower amounts. Results from 13C NMR spectroscopy and methylation analysis indicated the presence of branched arabinans, (1?>4)-and (1?>3,6)-linked galactans and hemicelluloses of the xylan and glucan types. The pectins differed in theirmolecular properties. HP exhibited a low molecular weight (M W = 45,000), whereas RP had a broad M W-distribution with M W= 360,000. Independently of these differences, the binding capacity of both pectins towards divalent cations was influenced by the DE only and increased in the order: Ca2+ < Cd2+ < Pb2+ < Cu2+.  相似文献   

18.
The aim of this research was to develop a quantitative treatment of the consequences of relatively slow initiation on M n and N (the number of molecules formed, Wp/M n , where Wp =weight of polymer formed) in living carbocationic polymerizations, particularly for the case of the incremental monomer addition (IMA) technique. This has been achieved by analysis of the effect of initiator efficiency (Ieff (%) = 100N/[I 0], where [I 0] = initiator input) on M n versus Wp , and N versus Wp plots. Three types of systems have been discerned: 1) Ieff equal to 100%; 2) Ieff constant but less than 100%; and 3) Ieff less than 100% but increasing with increasing number of monomer increments j by the IMA technique. Thus conditions can be found under which slowly initiating systems yield close to “ideal” product, i.e., one with M n = [M0 ]/[I0 ] and narrow molecular weight distribution (M w /M n ≈ 1.1). The corresponding equations and plots can be used to diagnose the mechanism. Subsequently, this quantitative analysis was used to describe a novel living system, trans‐2,5‐diacetoxy‐2,5‐dimethyl‐3‐hexene (DiOAcDMH6)/BCI3/isobutylene/CH3CI. This system produces linear t‐chlorine‐telechelic polyisobutylenes under homogeneous conditions. Surprisingly, cationation seems to be rate determining. This conclusion is illustrated by chemical equations.  相似文献   

19.
The polymerizable surfactant sodium 11-acrylamidoundecanoate (Na 11-AAU) was synthesized from acryloyl chloride and 11-aminounde-canoic acid. It had a low critical micelle concentration (CMC) of 4.3 × 10?4 mol/L. Polymerization of Na 11-AAU initiated by K2S2O8 was very fast in aqueous solution, with an activation energy of only 63.2 kJ/mol. The polymerization followed first-order kinetics with respect to Na 11-AAU and one-half order with respect to K2S2O8. The MW of poly(Na 11-AAU) was very high (1–2 million) but the MWD was rather narrow ( M w / M n = 1.45). Polymerization of Na 11-AAU in the micellar state may be responsible for the phenomena observed.  相似文献   

20.
The mechanism of polymerization of p-tert-butylstyrene (ptBuSt) initiated by the cumyl chloride/BCl3 initiating system in CH2Cl2 at -50°C has been investigated. At and below ~0.4 M ptBuSt, quasiliving polymerizations proceed, i.e., initiation is instantaneous, termination is absent or reversible, and chain transfer to monomer can be suppressed or eliminated. In the quasiliving range the M n versus [ptBuSt]0 plot is linear and passes through the origin, and a M w/M n decreases much below 2.0 with decreasing [ptBuSt]. GPC traces change from broad multimodal to narrow monomodal and the color of polymerization charges change from colorless to golden-yellow with decreasing [ptBuSt]. The effect of temperature jump subsequent to monomer addition has been examined; however, it does not explain the peculiar monomer concentration effect on the mechanism. Changes in the ionicity may be responsible for this phenomenon.  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