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 31 毫秒
1.
N.L Thomas  A.H Windle 《Polymer》1982,23(4):529-542
A theory is proposed to explain the transport behaviour of organic penetrants in glassy polymers in terms of two basic parameters: the diffusivity of the penetrant, D, and the viscous flow rate of the glassy polymer, 1η0. The rate controlling process for transport in these systems is considered to be diffusion of solvent down an activity gradient coupled with time-dependent mechanical deformation of the polymer glass in response to the swelling stress. The theory combines these two factors and is able to predict a wide range of observed transport phenomena from Fickian diffusion kinetics at one extreme to so-called Case II and Super-Case II behaviour at the other. The existence of a sharp front separating swollen and unpenetrated polymer is shown to result from the concentration dependence of the viscous flow rate.  相似文献   

2.
N. Kuwahara  M. Nakata  M. Kaneko 《Polymer》1973,14(9):415-419
Cloud-point curves for solutions of five polystyrene samples, including three well-fractionated polystyrenes, in cyclohexane have been examined near their critical points. Even for a solution of polystyrene characterized by MwMn<1.03, the critical point determined by the phase-volume method is generally situated on the right hand branch of the cloud-point curve. The precipitation threshold concentration is appreciably lower than the critical concentration, while the threshold temperature slightly deviates from the critical temperature. The agreement of the precipitation threshold point with the critical point has been found for a solution of polystyrene characterized by Mw=20×104 and MwMn<1.02 in cyclohexane. The η(φ) function derived from critical miscibility data is expressed by χ(φ) = 0.2798+67.50T+0.3070φ+0.2589φ2, which yields θ of 33.2°C and ψ1 of 0.22.  相似文献   

3.
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   

4.
A flow microcalorimeter designed to measure the heat of mixing of dilute polymer solutions is described. The instrument is sensitive to steady state heating rates of ~10 μJ/sec. Measurements of heats of mixing of solutions of differing concentrations of n-hexane and cyclohexane are reported and are compared with recommended data of McGlashan and Stoeckli. Values of:
K1=limV2→ 0
(H?1 ? H?01RTv22 are obtained for four polymer—solvent systems: polyisobutylene—benzene, 0.22; polystyrene (PS)—cyclohexane, 0.33; PS—n-butyl acetate, ?0.06 all at 25°C; and PS—toluene, ?0.05 at 40°C. Various theoretical calculations of second virial coefficients A2 made with use of the calorimetric data are compared with previously measured A2 for the first two mixtures.  相似文献   

5.
GPC data have been measured for a series of acetylated solvent refined lignite (SRL) asphaltenes and preasphaltenes and model compounds. Two structural parameters, the degree of aromatic condensation (HaruCar) and the molar hydrogen to carbon ratio (HC), were employed to correct the molecular weight of the samples for linear molecular size. For the model compounds, HaruCar was more effective, whereas the SRL materials were better adjusted by HC. The calibration standards deemed most suitable for determination of molecular weight of SRL by GPC are the SRL samples themselves.  相似文献   

6.
G.B. McKenna  L.J. Zapas 《Polymer》1983,24(11):1495-1501
The torsional behaviour of 1, 3 and 5 phr peroxide crosslinked natural rubber has been characterized over a range of strains from near the undistorted state (γ ≈ 0.017) to γ ≈ 1.0. Isochronal measurements of both torque and normal force were used to calculate values of the derivatives of the strain energy function W with respect to the first and second stretch invariants I1 and I2. In the course of our work we found that, contrary to many reports in the literature, ?W?I1 was affected significantly by the amount of crosslinking. Finally for the 1 phr peroxide crosslinked rubber it was found that, while ageing for 14 months at ambient conditions did not significantly affect the small-strain torsional modulus, G = 2(?W?I1 + ?W?I2), it did significantly affect the individual derivatives ?W?I1 and ?W?I2.  相似文献   

7.
A combination of steady-state and fluorescence decay techniques permits one to measure the dynamics of end-to-end cyclization of a polymer chain substituted at both ends with pyrene groups. In the limit of low concentration, the rate constant for cyclization, kcy, can be identified with the slowest relaxation rate τ1?1 of a Rouse—Zimm chain. Experiments are reported which allow kcy to be examined for two chain lengths of polystyrene substituted on both ends with pyrene groups. These chains have M?n = 9200 and 25 000 (M?wM?n ? 1.15). Added unlabelled polystyrene polymer [PS] causes k?cy to decrease in cyclohexane just above the θ-temperature, whereas in toluene, a good solvent, kcy is largely unaffected, even at [PS] concentrations of 50 wt%. These results are explained in terms of frictional effects—hydrodynamic screening—dominating in the poor solvent, whereas other factors tend to have offsetting effects in the good solvent.  相似文献   

8.
9.
Polymerization of ethylene and propylene was carried out with a soluble Cr(C17H35COO)3AlEt2Cl catalytic system in toluene. The system was found to be active only for ehtylene polymerization. From e.s.r. measurements of the catalytic system combined with some qualitative experiments, the active species was determined to be Cr2+. The reason why it should be incapable of polymerization of propylene was also discussed. The catalytic system showed some activity for ethylene-propylene copolymerization to give a random copolymer with a narrow molecular weight distribution.  相似文献   

10.
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   

11.
12.
Kock-Yee Law 《Polymer》1984,25(3):399-402
The effect of solvent vapour on the properties of vapour-swollen vinyl chloride-vinyl acetate (8317) copolymer has been studied by a fluorescence probe, a p-N,N-dialkylaminobenzylidenemalononitrile derivative (1). Results show that the fluorescence quantum yield (Фf) of 1 in vinyl chloride-vinyl acetate (8317) matrix decreases by a factor of ≈ 10, an indication of the increase in free volume or in polymer chain mobility, upon vapour swelling. The variations of Фf observed in various swollen matrices, which correlate only with the density of the swelling solvent, indicate that there is a profound vapour effect on the properties of swollen polymer. A density effect on the mobility of polymer chains in swollen polymer is proposed.  相似文献   

13.
Torsion and normal force measurements were made during single step stress relaxation experiments on a polymeric glass (PMMA). Isochronal data were analysed using an approach adapted from that developed by Penn and Kearsley1 (for incompressible elastic materials) to determine the derivatives ?W?I1, and ?W?I2 of the time dependent strain potential function. ?W?I1 and ?W?I2 are determined from existing solution to the torsion of an incompressible cylinder. A special solution to the torsion of a compressible cylinder is presented and it is shown that the values of ?W?I1 and ?W?I2 obtained using this solution to analyse the data do not differ greatly from those obtained using the incompressible solution. It is found from both solutions that ?W?I1 is negative and increases towards zero with increasing time and deformation while ?W?I2 is positive, greater in magnitude than ?W?I1 and decreases towards zero with increasing time and deformation. These results were unexpected and a full understanding of their meaning has yet to be reached.  相似文献   

14.
Raman spectra of sulfided Moγ-Al2O3 catalysts were obtained using in situ techniques for two sulfiding methods. For samples sulfided by 10% H2SH2 at 400 °C, MoS2 structures were observed. A stepwise sulfiding using 10% H2SH2, with spectra recorded at 150, 250, and 350 °C, resulted in observation of molybdenum oxysulfide, reduced molybdate, and surface “MoS2” phases. Reexposure of these samples to air led to radical modification of the oxysulfide structures as well as transformation of some sulfide phases. A model incorporating terminal and bridging MoS bonding and anion vacancies is proposed. This model is based on the conversion of isolated and aggregated molybdate and MoO3 species to oxysulfide and reduced molybdenum phases. Conversion of reduced molybdenum phases to sulfides is observed to be slow.  相似文献   

15.
Ivo Lang  Milan Hájek 《Fuel》1985,64(11):1630-1631
Brown and Ladner's parameters have been compared with those of Williams for the same conditions. The combination of experimental data from 1H n.m.r. and 13C n.m.r. spectra gives the value of the Brown and Ladner's parameter HaruCar. A comparison of the HaruCar values was performed on a group of different coal oil samples.  相似文献   

16.
The formation of the Pb(Mg13Nb23)O3 and Pb (Fe12Nb12)O3 phases with a perovskite type structure is directly dependent on the reactivity of magnesium and ferric oxides to other phases belonging to the binary system PbO  Nb2O5. Moreover, it is shown that the ceramic process influences the proportion of perovskite phases in comparison with parasite phases and also the densification of the samples. The optimization of the ceramic process allows to obtain a pure Pb(Fe12Nb12O3 phase, but as far as Pb(Mg13Nb23)O3 is concerned, a parasite phase is never entirely eliminated.  相似文献   

17.
A study of CO hydrogenation over PdSiO2 and PdLa2O3 has been carried out for the purpose of identifying the effects of Pd dispersion, Pd morphology, and support composition on the catalytic activity of supported Pd. The specific activity of each catalyst for methanol and methane synthesis was determined from microreactor studies carried out at a fixed set of reaction conditions. Palladium dispersion was measured by H2O2 titration, and the morphology of the Pd crystallites, as expressed by the distribution of Pd(100) and Pd(111) planes, was determined from in situ infrared spectra of adsorbed CO. The crystallite morphology of the PdSiO2 catalysts is the same, independent of Pd weight loading: 90% of the surface is comprised of Pd(100) planes and 10% of the surface is comprised of Pd(111) planes. By contrast, the crystallite morphology of the PdLa2O3 catalysts changes with Pd loading. Primarily Pd(100) planes are exposed at low-weight loadings while Pd(111) planes are exposed at high-weight loadings. The Pd dispersion has little effect on the methanol turnover frequency over both PdSiO2 and PdLa2O3, for dispersions between 10 and 20%. On the other hand, the methane turnover frequency is independent of Pd dispersion over PdSiO2, but increases with decreasing dispersion over PdLa2O3. It is further observed that the Pd morphology influences the specific activity of PdLa2O3 for methanol synthesis: Pd(100) is nearly threefold more active than Pd(111). For a fixed morphology, the specific methanol synthesis activity of PdLa2O3 is a factor of 7.5 greater than that of PdSiO2.  相似文献   

18.
High resolution neutron scattering experiments have been used to observe the diffusive motion of low molecular weight linear and cyclic poly(dimethyl siloxane) molecules in dilute solution in deuterated benzene. Diffusion coefficients (D) and hydrodynamic radii (RH) have been compared with values obtained by light scattering for higher molecular weight samples and with radii of gyration (Rg) obtained by small-angle neutron scattering. While the ratio DringDchain is close to the predicted value of 0.85, the ratio RgRH falls below the theoretical value for both ring and chain molecules. The scattering curves show effects arising from both centre of mass diffusion and internal molecular motion, and the observed inverse correlation times are compared with calculated behaviour as a function of scattering vector, Q.  相似文献   

19.
Klaus J. Hüttinger  P. Schleicher 《Fuel》1981,60(11):1005-1012
The catalysis of hydrogasification of carbon by Fe, Co and Ni was studied using a special petroleum coke with extremely low reactivity. The kinetics were studied with impregnated coke in a fixed-bed flow reactor between 1133 and 1235 K and up to 2 MPa, yielding the following rate equation: ?rH2 = k(CH2?CH2e)(1 + Kads · CH2)2Apparent activation energies and heats of adsorption are: Fe, 152 and ?92 kJ mol?1; Co, 201 and ?82 kJ mol?1 Ni, 165 and ?50 kJ mol?1. These studies with impregnated coke as well as further gasification experiments with cokes heat-treated after impregnation with metal salts up to 2273 K confirmed a spillover mechanism and excluded any influence of electronic interactions between carbon and the catalyst metals.  相似文献   

20.
The melting behaviour of drawn poly(ethylene terephthalate) bristles has been studied by means of differential scanning calorimetry. In addition the wide-angle X-ray diffraction pattern were analysed. For comparison some of the experiments were also carried out with undrawn samples. The differences in the melting curves of drawn and undrawn PET originate from the different crystallization kinetics. The density defect (?idc ? ?1c) between the ideal crystal density ?idc and the effective density ?1c of the crystalline layers is a result of lattice vacancies introduced by the grain boundaries of the mosaic blocks. The relatively low ultimate crystallinity of PET is supposed to be caused by the hindrance of crystal growth of fibre direction during isothermal crystallization.  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