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 937 毫秒
1.
Wyn Brown  Peter Stilbs 《Polymer》1983,24(2):188-192
Transport in ternary polymer1, polymer2, solvent systems has been investigated using an n.m.r. spin-echo technique. The dependence of the self-diffusion coefficient of poly(ethylene oxide) polymers on the concentration and molecular size of dextran in aqueous solution has been measured. Monodisperse poly(ethylene oxide) fractions (M?w=7.3×104, 2.8·105 and 1.2·106) and dextrans (M?w=2·104, 1·105 and 5·105) have been employed over a range of concentration up to the miscibility limit in each system. It is found that when the molecular size of the diffusant is commensurate with or exceeds that of the matrix polymer, a relationship of the form: (DD0)PEO=exp?k(C[η]) is applicable, where C[η] refers to the dextran component and is considered to describe the extent of coil overlap in concentrated solution. (DD0) is independent of the molecular size of the poly(ethylene oxide), at least in the range studied (Mw<300 000).  相似文献   

2.
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   

3.
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   

4.
The viscosity of heparin was examined in pure glycerol, formamide and in mixtures of the two solvents. This study confirms that the equation ηspc = [η] [1 + k(c)12], where [η] is the shielded intrinsic viscosity, adequately describes the concentration dependence of the reduced viscosity of the polyelectrolyte, heparin. The [η] linearly increases with the dielectric constant of the solvent.  相似文献   

5.
E. Straube 《Polymer》1985,26(1):105-108
A polymer chain consisting of Nr segments with a repulsive interaction (binary cluster integral βr) and Na ? Nr segments with a stronger, attractive and pairwise saturable interaction (βa), which is at the averaged θ-point N2rβr + N2aβa = 0 deviations from the predictions of the two parameter theory: α2R ? 1 ~ δzr < 0 and A2δzr > 0 with δzr ~ βr(NaNr)12. It is shown that the deviations from the universal behaviour are due to the existence of an intermediate length scale NaNr.  相似文献   

6.
B. Nyström  J. Roots  R. Bergman 《Polymer》1979,20(2):157-161
Sedimentation velocity measurements on polystyrene (M = 110 000) in cyclopentane over an extended concentration region and from 5°C (close to the upper critical solution temperature) to 40°C are reported. The concentration dependence parameter (ks)w increases from 2 to 5°C to 27 at 40°C. For all temperatures except 5°C, s0s vs. w[η]w shows an upward curvature at w[η]w ≈ 1; at 5°C, on the other hand, s0s is independent of concentration over the region considered. Furthermore, measurements have also been performed at 20°C (θ-conditions) over a large concentration interval for the molecular weights M = 20 400, 390 000 and 950 000. The parameters s0 and (ks)w were both found to be proportional to M?1/2w. In the ‘hydrodynamically normalized’ plot s0s vs. w[η]w the sedimentation behaviour can approximately be represented by a single curve for all the molecular weights.  相似文献   

7.
The kinetics of polymerization of the symmetrical non-conjugated divinyl monomer N,N′-methylenebisacrylamide has been studied using CeIV-thiourea redox system as initiator. The rate of polymerization, Rp is proportional to [CeIV]12, [thiourea]12 and [monomer]32. A cyclopolymerization mechanism fits in with the experimental results.  相似文献   

8.
The reaction of an equimolar mixture of SrCO3 and GeO2 proceeds in five stages, [1]–[5]. The overall reaction of [1], [2], and [3]
2SrCO3+GeO2Sr2GeO4+2CO2[1]SrCO3+4GeO2SrGe4O9+CO2[2]SrCO3+GeO2β-SrGeO3+CO2[3]β-SrGeO3transformationα-SrGeO3[4]3Sr2GeO4+SrGe4O9→7(α-SrGeO3)[5]
is best described by the Jander equation, the apparent activation energy being 47.6 kcal mol?1 irrespective of the ball-milling time. β-Strontium metagermanate is formed directly at lower temperatures from amorphous material prepared by the simultaneous hydrolysis of strontium and germanium alkoxides. Kinetic studies of β-SrGeO3 formation and β- → α-SrGeO3 transformation are carried out by means of X-ray diffraction.  相似文献   

9.
P. Törmälä  G. Weber 《Polymer》1978,19(9):1026-1030
The tumbling of five nitroxide spin probes (molecular weights between 172–486 g/mol) in a standard unfractionated polyisobutylene [M?v = (1.26 ± 0.18) × 106g/mol] has been studied by means of the electron spin resonance (e.s.r.) technique. The temperature at which the separation of the outermost peaks of the e.s.r. spectrum is 50 G (T50G) attained a limiting value T50G = 330K at probe MW = 332 g/mol. This temperature coincided with the temperature of the loss maximum of the merged glass transition (Tg) and segmental relaxations at the corresponding frequency (3 × 107 Hz). A literature survey indicated that an analogous situation exists in the case of poly(vinylidene fluoride) and polyamide-6,10 while T50G values of poly(2,6-dimethyl phenylene oxide) and polycarbonate are correlated only to segmental relaxations of polymer chains. It is concluded that the equation:
T50G = Tg[1 + (exp Tg/Tc)?1]
describes generally the temperature shift between glass transitions at low and high frequencies and can be applied to determine experimentally low frequency Tg values from T50G values if Tg and T < Tg relaxations (if any are present) are already merged at this temperature.  相似文献   

10.
Hydrotreatment of spent oil distillate was carried out on a commercial Ni-Mo-alumina catalyst in the temperature range 260–340 °C, with a liquid hourly space velocity (LHSV) of 0.7–2.0 h?1, pressure of 4.5 MPa and H2oil ratio of 300 NL L?1 (normal litre of H2 per litre of feedstock). U.v. spectra of hydrogenated and original spent oil distillates (measured in normal hexane) gave a band with a maximum at 230 nm. The change in absorbance at three selected wavelengths for original oil distillate and hydrotreated oil at different operating conditions was taken as a guide for the determination of hydrogenation reaction rates (including partial saturation of aromatics and sulphur compound hydrogenolysis). The rate constants of hydrogenation reactions (k) using a second-order equation and a model of two parallel first-order reactions (k1 and k2) were calculated. Finally, the apparent activation energy (Ea), enthalpy of activation (ΔH1) and entropy (ΔS1) were calculated based on the values of k, k1, and k2. The calculated values of Ea based on k, k1 and k2 were 81.479, 71.188 and 62.882 kJ mol?1, respectively. The values of ΔH1 based on the same rate constants were 76.670, 66.564 and 58.433 kJ mol?1, while the values of ΔS1 were ?117.150, ?133.779 and ?150.823 J mol?1 K?1, respectively.  相似文献   

11.
Extremely high molecular weight polystyrenes with a M?w in the range 10.8 × 106 to 2.2 × 107 were prepared by emulsion polymerization initiated with a heterogeneous initiator at 30°C, which has a ‘living character’. Samples of polystyrene were characterized by light scattering and viscometry in toluene and benzene at 25°C, and in θ-solvent cyclohexane at 34.8°C. Also determined were the relationships of mean-square radius of gyration 〈s2〉 (m2) and the second virial coefficient A2 (m3 mol kg?2) on the molecular weight, which for toluene and benzene are described in equations: Toluene (25°C) 〈s2〉=1.59 × 10?23M?w1.23; A2=4.79 × 10?3M?w?0.63; Benzene (25°C) 〈s2〉=1.23 × 10?22M?w1.20; A2=2.59 × 10?3M?w?0.59. The parameters in the Mark-Houwink-Sakurada equation were established, for extremely high molecular weight polystyrene in toluene and in benzene, at 25°C into the form giving for [η] (m3kg?1): [η] = 8.52 × 10?5M?w0.61; [η] = 1.47 × 10?4M?w0.56. The mentioned relations, as well as the obtained values of Flory parameter ?0 and of ratio [η]M?w0.5 were compared with solution properties of high molecular weight polystyrene with narrow molecular weight distribution prepared by anionic polymerization by Fukuda et al.  相似文献   

12.
A Monte Carlo method has been devised for calculating the conformation-dependent properties of cyclic poly(dimethyl siloxanes) (PDMS), using Flory, Crescenzi and Mark's rotational isomeric state model. Calculated values of the mean-square radii of gyration 〈s2r〉 of ring molecules unperturbed by excluded volume effects and containing 8–100 skeletal atoms are compared with the 〈s2l〉 values for the corresponding unperturbed chain molecules. Exact enumeration methods were also employed for rings [(CH3)2SiO]w2 with w ? 24 and the results found to be in close agreement with those obtained by the Monte Carlo method. The ratio 〈s2l〈s2r was found to attain limiting values close to 2.0 for w > 30, in agreement with theoretical predictions.  相似文献   

13.
C.H Bamford  E Schofield 《Polymer》1983,24(4):433-438
The different kinetic features associated with polymerizations in which retardation arises through degradative transfer and degradative addition processes are considered. Reasons are given for believing that in reactions retarded by degradative transfer neglect of the interaction between ‘inactive’ radicals is justifiable, while with degradative addition this is not so. A kinetic treatment of the latter is developed in which all three termination reactions between propagating and inactive radicals are assumed to be diffusion-controlled with a single rate coefficient and equations are derived which permit from experimental data on rates of polymerization estimation of the definitive kinetic parameters kpk′?12t, kpmk′?12t and kfmk′?12t (kp, kt, kpm, kfm are the rate coefficients of propagation, termination, reinitiation and degradative additions, respectively). The kinetic equations are satisfactorily consistent with ew experimental data on the polymerization of 1-vinylimidazole in ethanol solution at 70°C, for which there is strong evidence for the occurrence of degradative addition. A modified procedure for processing data for polymerizations with degradative transfer is put forward which is convenient for estimating the kinetic parameters and reveals in a simple manner the importance of re-initiation. It is suggested that this treatment could be generally useful in the early stages of the study of a retarded polymerization.  相似文献   

14.
Shaul M. Aharoni 《Polymer》1980,21(12):1413-1422
When plotted against concentration V02, the viscosity η0 curve of isotropic solutions of lyotropic nematic mesomorphic polymers increases according to:
η00[[η]V02+(π4)[η]2(V02)2+K2(lnx)2[η]3(V02)3+…]
in which η0 is the solvent viscosity, K2 a numerical constant, x? is the average molecular axial ratio and [η] the intrinsic viscosity as defined by:
[η]=2x2451ln 2x?1.84+3ln 2x?0.61 (+1415)
At the concentration v12 an anisotropic phase appears and the viscosity curve shows a decrease in slope followed by a change in direction at a peak viscosity ηp at vp2. Upon further increase in v02 the system undergoes a phase inversion and finally turns fully anisotropic at vA2. In the biphasic interval the viscosity is described by:
η0mat 1+(5λ+2)2(λ+1)Vinc
where λ = ηincηmat, the ratio of the inclusions viscosity to the matrix viscosity, and vinc the volume fraction of the inclusions in the system. It must be emphasized that the molecular weight of the polymer in both phases changes continuously with concentration, resulting in commensurate changes in ηince, ηmat and their ratio λ. In the anisotropic region the viscosity first decreases moderately and then increases precipitously with v02, according to:
η00[2x290S(5ln 2x?1.8+6ln 2x?0.61)+2]V021-V25°VE
in which S is an order parameter and V25, VE are volumes swept by the orbits of the flowing rodlike macromolecules. The equations give results in good qualitative and fair quantitative agreement with experimental data in the literature.  相似文献   

15.
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   

16.
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   

17.
18.
The relaxational and retardational properties of poly(propylene glycol) liquids, of nominal molecular weights 400 and 4000, are described. The viscoelastic behaviour of each liquid has been determined over a wide temperature range, using high frequency shear wave techniques operating at 30 and 454 MHz. It is found that the complex compliance J1(jω) is described in terms of the viscosity ν, the limiting high frequency compliance J, the retardational compliance Jr and a characteristic retardation time тr by:
J1(jω)=J+1jωη+jr(1+jωτr)β
where β is a parameter of the retardation time distribution.For the lower molecular weight liquid, JrJ = 17.4, β = 0.45 and тrтm increases with decreasing temperature, reaching a limit of 170 near 0°C. This liquid shows no evidence of polymeric behaviour.For the other, JrJ, β = 0.76, тrтm = 15.4 and is constant over the temperature range investigated. The main difference between the two liquids appears as an additional retardational or relaxational process for the higher molecular weight material which occurs in the initial or low frequency part of the relaxation region. This process is characterized by a single time, but with relaxation time 17 and a stiffness 7 times the values calculated for the first Rouse mode of polymer chain motion.  相似文献   

19.
20.
Diffusion of methane from three coals ranging in rank from anthracite to HVA bituminous has been studied at initial methane pressures up to about 2.76 MPa (400 psi). Unsteady-state diffusion conditions existed, the methane pressure within the coal particle decreasing with time while the methane pressure outside the particles remained at atmospheric. The diffusion parameter D12r0 increased with increasing methane concentration at high values of methane sorption. Diffusion was activated but the exact magnitude of the activation energy is uncertain owing to the suspected contribution of the heat of sorption to the temperature coefficient. D12r0 increased with decreasing particle size of coal studied, but r0 is clearly less than the particle radius.  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