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 15 毫秒
1.
2.
The apparent diffusion coefficients for Ti, V, Cr, Nb, Mo and Hf as carbides and for elementary Fe, Ni and Cu in electro graphite have been determined by means of an electron-microprobe analyzer. These pseudo diffusion coefficients were found to vary with the heat treatment time. However, after one hour these remain constant and follow the Arrhenius type of relation D = D0exp(?Q/RT). The activation energy Q was nearly constant for the metals investigated. An attempt was made to correlate the frequency factor D0 with the heat of formation ΔH?298 of the corresponding carbides. A plot of log D0vsΔHf yielded two straight lines, one for the negative ΔH?, the other for positive ΔH?. This method was satisfactorily applied to predict the diffusion coefficients of Zr, Sb and Bi.  相似文献   

3.
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   

4.
Walter Burchard 《Polymer》1979,20(5):577-581
The influence of a size distribution on the angular dependence of the quasielastically scattered light is studied for (i) large hard spherical particles and (ii) large flexible chain molecules. For the spherical particles the angular dependence is shown to depend solely on the size distribution and the particle scattering factor. Combination of conventional elastic light scattering with quasielastic light scattering allows the determination of the z-average radii moments rn?z (n = ?1, 1, 2, …) which define the size distribution. Flexible chains — linear and branched ones — show in any case a linear dependence of the apparent diffusion constant Dapp = Гq2 on q = (λ) sin θ2, when q becomes large. This behaviour represents the flexibility of the spring-bead model with strong hydrodynamic interaction. The initial part on the other hand form a straight line when Dapp is plotted against q2. The intercept of this straight line is the z-average diffusion constant while the slope is proportional to the z-average mean square radius of gyration. Thus, the polydispersity can be estimated from Dz and 〈S2z while the asymptote at large q-values is determined by the internal flexibility of the molecule.  相似文献   

5.
6.
Using the Maxwell–Stefan theory for diffusion we derive a simple formula to relate the tracer (i.e. self) diffusivity D1 and Maxwell–Stefan (MS), or jump, diffusivity D. The presence of the interchange coefficient Dij in the MS formulation causes the self diffusivity to be lower than the jump diffusivity. Assuming the interchange coefficient to be given by D/F we derive:D1=D1+Fθwhere F is a factor to take account of topology effects within the zeolite matrix.The validity of the MS formulation is established by performing kinetic Monte Carlo simulations for diffusion of methane, perfluoromethane, 2-methylhexane and iso-butane in silicalite. Furthermore, it is shown that the exchange coefficient Dij is a quantification of correlation effects during the hopping of molecules. For iso-butane, the isotherm inflection leads to a sharp inflection in the diffusion behaviour. The influence of molecular repulsive forces on the loading dependence of the jump and self-diffusivities is also discussed with the aid of published Molecular Dynamics simulations for methane.  相似文献   

7.
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   

8.
Emulsion polymerization of vinylacetate leads to branched polymers which at high monomer conversions form microgels of the shape and size of the latex particles. Quasielastic light scattering measurements from samples in the pre-gel state give at small q2 a linear angular dependence of Dapp = Гq2 which resembles that of randomly branched chain molecules, where Г is the decay constant of the time correlation function. Extrapolation of Dapp towards zero scattering angle yields the translational diffusion constant Dz. The diffusion constant follows the molecular weight dependence Dz = 9.78 10?5Mw?0.478. The diffusion constant of the microgels, i.e. at molecular weights Mw > 14 106, remains constant because of the finite and constant size of the latex particles. The coefficients kf and kD in the concentration dependence of the frictional and diffusion coefficients are related according to the equation kD = kf ? 2A2Mw ? v? where A2 is the second virial coefficient and v? the partial specific volume of the particle. The coefficient kf is calculated from the experimentally determined quantities kD, A2 and Mw, and the result is compared with the theory by Pyun and Fixman. Accordingly the branched coils in the pre-gel state resemble soft spheres, but the microgels behave more like spheres of some rigidity.  相似文献   

9.
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   

10.
G.B. McKenna  K.L. Ngai  D.J. Plazek 《Polymer》1985,26(11):1651-1653
Within the context of a generalized coupling model we can support the hypothesis that, while the mode of relaxation for self diffusion (D) and shear flow (η) are the same, the entanglement interactions are different. We assume that there are two distinct coupling parameters nD and nη for self diffusion and shear flow respectively. The model predicts the molecular weight and temperature dependences to be scaled by the relevant coupling parameters as:
η∝[M2exp(Ea/kT)]1(1?nη)and D∝M[M2exp(Ea/kT)]?1(1?nD)
for melts with Arrhenius temperature dependences. We have found that nn=0.43 and 0.42 for polyethylene (PE) and hydrogenated polybutadiene (HPB) which scale η as M3.5 and M3.4. Also the apparent flow activation energies E1a of 6.35 kcal mole?1 for PE and 7.2 kcal mol?1 for HPB scale to primitive activation energies Ea of 3.6 and 4.2 kcal mole?1 for PE and HPB respectively. On the other hand the M?2 dependence of D results in nD=1/3. Then the reported activation energies for self-diffusion in PE and HPB of 5.49 and 6.2 kcal mole?1 scale to primitive activation energies of 3.7 and 4.1 kcal mole?1, respectively.  相似文献   

11.
W. Brown  R.M. Johnsen 《Polymer》1981,22(2):185-189
Using a novel sorption technique, the diffusion of some series of solutes in polyacrylamide gels has been investigated with regard to: (a) molecular size of solute; (b) concentration of solute and gel polymer; and (c) temperature. The approach used also yields the partition coefficient pertaining to sorption equilibrium. The ratio, DDo, where Do refers to diffusion in the pure solvent, is found to reflect in part the characteristic interactions between solute and gel polymer. The temperature results indicate that the apparent activation energy for solute diffusion is approximately independent of the polymeric component for dilute gels.  相似文献   

12.
Measurements of the real portion of the dielectric constant of coal-water slurries at 750MHz are reported. Slurries of 40–65wt% of either Illinois No. 6, Utah bituminous or Texas lignite coals over the temperature range 11–71 °C were studied. The dielectric constant was independent of coal type and particle size within experimental error. The measured values of dielectric constants agreed with those predicted by the Looyenga equation: ?′ = [?′132 + φ(?′131??′132)]3 where: ?′, the mean value of the dielectric constant; ?′1, the dielectric constant (3.8) of the dry coal; ?′2, the dielectric constant of water; φ, the volume fraction of coal.  相似文献   

13.
A.R. Greenberg  R.P. Kusy 《Polymer》1984,25(7):927-934
The applicability of the Gibbs—DiMarzio (G—DM) theory of the glass transition (Tg) is quantitatively evaluated for PS, PVC, PαMS and PMMA. The analysis was conducted under the assumption that both the inter-/intramolecular energy ratio (r) and the effective chain segment density (n) remain constant while the fractional free volume at Tg(V0) varies as a function of the reciprocal degree of polymerization (103P?). Based upon reduced parametric plots of TgTg∞versus103P?, the results showed that the G-DM equations were satisfactory for PS and PVC but unsuccessful in the cases of PαMS and PMMA. For the former cases the analysis indicated that when 0.015 ? V0 ? 0.045 optimum agreement occurred at n=1.80, r=10.5 and n=1.36, r=0.95, respectively. Although potential n, r values were obtained for PαMS when the allowable V0 range was expanded to 0.010–0.050, none of these combinations satisfied all of the analytical requirements. No agreement for the PMMA data sets could be obtained even when this less stringent V0 criterion was adopted. Attempts to improve this situation by incorporating ‘beads’ and ‘flexes’ into the statistical mechanical equations are also considered.  相似文献   

14.
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   

15.
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   

16.
Thermoplastic behaviour of a Pittsburgh seam hvA coal (PSOC1099) was characterized by the use of a high-pressure microdilatometer. Phenomena such as softening, swelling, final resolidification, and the temperatures at which they occur were measured as functions of heating rate (25 ° and 65 °C min?1), particle size (= 75 μm and 250 × 425 μm), gaseous atmosphere (N2, H2, COH2) and applied gas pressure (atmospheric to 2.8 M Pa). The results obtained illustrate several important aspects of thermoplastic properties of this coal under the conditions utilized. It is observed that pressure alone can play a major role in determining its overall thermoplastic behaviour. Compared to that at atmospheric pressure, swelling is significantly reduced at 2.8 MPa of pressure for any given heating rate or particle size. In these experiments, the chemical composition of the gaseous atmospheres (COH2, H2 and N2) does not appear to alter significantly the plastic phenomena at any given pressure. Increasing the heating rate or decreasing the particle size results in increased swelling at all applied pressures and atmospheres.  相似文献   

17.
J. Ehrlich  S.S. Stivala 《Polymer》1974,15(4):204-210
A bovine heparin fraction was examined by sedimentation analysis and intrinsic viscosity measurements as a function of ionic strength in the range of 0·1 to 1·0 M, and at pH 2·5 and 6·0. The following experimental parameters were obtained: M, S020,W, D020,W, V?, and [η]. Other physical parameters were calculated based on a random coil model (supported by the theory of Mandelkern and Flory) e.g., (r?2)12, (s?2)12. Similar studies were made on a heparin sample as a function of desulphation as resulting from graded mild hydrolysis. Since desulphation is accompanied by decreasing anticoagulant activity of heparin, the latter was correlated with various calculated and measured physical parameters. Significantly (r?2)12 decreases with decreased desulphation and therefore decreased biological activity.  相似文献   

18.
Dense tungsten hemicarbide specimens were used to study the bulk and grain boundary diffusion of 14C into W2C at temperatures of 1200 to 2000°C. The bulk diffusion coefficient is given by:
Dv =18.3exp?91,500RTcm2s?1
The grain boundary diffusion coefficient is represented by the expression:
PGB =1.8 10?4exp?68,880RTcm2s?1
A comparison is give with preceding results on other carbides.  相似文献   

19.
Souheng Wu 《Polymer》1985,26(12):1855-1863
The effects of rubber particle size and rubber-matrix adhesion on notched impact toughness of nylon-rubber blends are analysed. A sharp tough-brittle transition is found to occur at a critical particle size, when the rubber volume fraction and rubber-matrix adhesion are held constant. The critical particle size increases with increasing rubber volume fraction, given by dc = Tc{(πr)13 ? 1}?1, dc is the critical particle diameter, Tc the critical interparticle distance, and ør the rubber volume fraction. The critical interparticle distance is a material property of the matrix, independent of rubber volume fraction and particle size. Thus, the general condition for toughening is that the interparticle distance must be smaller than the critical value. Van der Waals attraction gives sufficient adhesion for toughening. Interfacial chemical bonding is not necessary. Even if there is interfacial chemical bonding, a polymer-rubber blend will still be brittle, if the interparticle distance is greater than the critical value. The minimum adhesion required is about 1000 J m?2, typical for van der Waals adhesion. In contrast, chemical adhesion is typically 8000 J m?2. The present criterion for toughening is proposed to be valid for all polymer—rubber blends which dissipate the impact energy mainly by increased matrix yielding.  相似文献   

20.
The oxidation process in irradiated polyethylene was investigated by means of the e.s.r. method, and diffusion of oxygen molecules into the crystalline region of polyethylene was studied in detail. Computer simulation was carried out in order to determine various parameters of the process including diffusion constant. In the course of the simulation, the crystallite was assumed to have a plate-shaped form and the diffusion constant was considered to be larger at the region near the surface of the crystallite, Df, and to be smaller at the inner region of the crystallite, Ds. Making these assumptions in the simulation gave a much better result than the assumption based on a sphere-shaped crystallite. DfDs was found to be about 2 for thicker crystallite and it was 4 ~ 5 for thinner crystallite. Diffusion constants at various temperatures were determined, and the order of magnitude of diffusion constants in the crystalline region was found to be 10?16 cm2/sec at about 320K both for thicker and thinner crystallites. The activation energy of diffusion of oxygen in the crystallites was found to be 32 kcal/mol.  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