首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 187 毫秒
1.
The swelling features of gelatine gels in water (good solvent) were studied as a function of thermodynamic conditions of sol—gel transition and ripening. It is shown that the degree of equilibrium swelling Qe varies with the volume fraction of the polymer in a casting solution φo in accordance with the prediction of the classic theory: Qe φo−0.4. Qc, as a function of the gelation temperature Tg, the ripening time tr and φo, can be rescaled and described by the single empirical equation: Qe Tgx tryφo−0.4, where x = 0.1, y = 0.15 for wet gels and X = −0.5, y = 0.04 for dried gels. The kinetics of macroscopic swelling is described by the equation of Peters and Candau, with values of collective diffusion coefficients being in good agreement with values obtained by other workers via photon correlation spectroscopy.  相似文献   

2.
The electrosorption properties of p-norborn-2-yl phenolate ions in alkaline solutions were investigated by ac polarographic and electrocapillary measurements.

Two adsorption regions were found. At low bulk surfactant concentrations the adsorption at the positively charged electrode (−0.2 E −0.6 V) is predominant while at higher surfactant concentrations the adsorption at the negatively charged electrode (−0.6 E −1.0 V) is more pronounced. At E = −0.40 V the adsorption parameters were determined (a ≈ 2; ΔG°A = −32.5 ± 1 kJ mol−1. Between −0.6 E −1.0 V one potential of maximum adsorption for all concentrations does not exist and therefore the adsorption parameters could not be calculated.

At E = −0.40 V progressive two-dimensional nucleation with a nucleation order of 3 was observed which corresponds well with the high attraction constant.

The electrode reaction S2O2−8 + 2e → 2 SO2−4 is inhibited by norborn-2-yl phenolate ions in the potential range −0.2 E −0.6 V. In the second potential range of capacity decrease the electrode process is much less retarded. At E = −0.40 V, in a similar manner as described for neutral molecules, a linear dependence of the log ks (ks apparent rate constant) on ln cA and π (π = surface film pressure), respectively, has been found.  相似文献   


3.
Ralph J. Tyler 《Fuel》1980,59(4):218-226
The devolatilization behaviour of ten bituminous coals was investigated under rapid heating conditions using a small-scale fluidized-bed pyrolyser. The pyrolyser operated continuously, coal particles being injected at a rate of 1–3 g h?1 directly into a heated bed of sand fluidized by nitrogen. Yields of tar, C1–C3 hydrocarbon gases, and total volatile-matter and an agglomeration index are reported for all coals. Maximum tar yields were obtained at about 600 °C and were always substantially higher than those from the Gray-King assay. Total volatile-matter yields were also substantially higher than the proximate analysis values. The maximum tar yields appear to be directly proportional to the coal atomic HC ratio. The elemental analysis of the tar is strongly dependent on pyrolysis temperature. The tar atomic HC ratio is proportional to that of the parent coal. The effect on the devolatilization behaviour of two coals produced by changes in the pyrolyser atmosphere and the nature of the fluidized-bed material were also investigated. Substituting an atmosphere of hydrogen, helium, carbon dioxide or steam for nitrogen, has no effect on tar yield and, with one exception, little effect on the hydrocarbon gas yields. In the presence of hydrogen the yield of methane was increased at temperatures above 600 °C. Tar yields were significantly reduced on substituting petroleum coke for sand as the fluid-bed material. A fluidized bed of active char virtually eliminated the tar yield.  相似文献   

4.
New aspects of the defect diffusion model (DDM) are presented. First, it is shown that if the correlation volume exhibits two dimensional scaling, e.g. grows in two dimensions more rapidly than in a third orthogonal direction, the standard Vogel relation is obtained. Second, it is pointed out that, independent of the dimensionality, ∂ ln σ/∂P should be proportional to ∂ ln σ/∂ ln T where the proportionality constant is −∂ ln Tc/∂P. It is shown that both of these results are consistent with the temperature and pressure variation of the electrical conductivity for 20:1 PPG:LiCF3SO3 below about 1.3 times the glass transition temperature, Tg. Finally, the DDM is compared with the free volume theory (FVT) of Dlubek et al. who have reconsidered “traditional” FVT and presented a formalism where the free volume is not proportional to the macroscopic volume. One aspect of the new FVT, which is difficult to understand is that the compressibility of the occupied volume is larger than the compressibility of the free volume.  相似文献   

5.
The glass transition of thermoplastics of different polydispersity and thermosets of different network structure has been studied by conventional differential scanning calorimetry (DSC) and temperature modulated DSC (TMDSC). The cooling rate dependence of the thermal glass transition temperature Tg measured by DSC, and the frequency dependence of the dynamic glass transition temperature T measured by TMDSC have been investigated. The relation between the cooling rate and the frequency necessary to achieve the same glass transition temperature has been quantified in terms of a logarithmic difference Δ=log10[|q|]−log10(ω), where |q| is the absolute value of the cooling rate in K s−1 and ω is the angular frequency in rad s−1 necessary to obtain Tg(q)=T(ω). The values of Δ obtained for various polymers at a modulation period of 120 s (frequency of 8.3 mHz) are between 0.14 and 0.81. These values agree reasonably well with the theoretical prediction [Hutchinson JM, Montserrat S. Thermochim Acta 2001;377:63 [6]] based on the model of Tool–Narayanaswamy–Moynihan with a distribution of relaxation times. The results are discussed and compared with those obtained by other authors in polymeric and other glass-forming systems.  相似文献   

6.
The binary adsorption characteristics of methane and ethane on dry coal to 40 atm pressure have been calculated from pure-component isotherms. In some coal seams, pressures exceeding 40 atm have been recorded and the methane sampled from the virgin coal often shows a few percent of ethane. The binary adsorption characteristics were calculated by employing the ideal adsorbed solution theory of Myers and Prausnitz, and experimentally-determined (Type I) pure gas isotherms at 0, 30 and 50 °C. The coal used in this investigation was high-volatile ‘A’ bituminous (hvab) from the Pennsylvania Pittsburgh seam. Gas nonideality was accounted for by replacing pressure with fugacity. Adsorption of methane on dry coal is purely physical; the isosteric heat of adsorption does not exceed 2.4 kcal/mol* at 30 °C on the above coal. Isobars on the resulting binary equilibrium diagram exhibited an unexpected phenomenon of intersecting each other which might be attributable to the above nonideality considerations. The region of a few percent of ethane, which is of practical importance from the viewpoint of coal seams, was expanded and reduced to an equation: V(CH4) = −21.52 + 7.18(VF) + 16.88(VF)2 −0.395(P) − 0.00661(P)2 + 0.824(T) − 0.00030(T)2 + 0.928(VF)(P) − 0.858(VF)(T). V(Total) = 25.9 − 23.6(VF) + 0.655(P) − 0.00875(P)2 − 0.795(T) + 0.743(VF)(T) where V(CH4) and V(Total) = cm3(STP)CH4 and total gas respectively adsorbed per g dry coal; VF = vol. fraction of methane as analysed at 1 atm (0.94 VF 1.0); P = seam pressure, atm (0 P 40); T=seam temperature, °C(−10 T 50).  相似文献   

7.
Edmund A. Di Marzio 《Polymer》1990,31(12):2294-2298
The entropy theory of glasses is used to derive the glass temperature, Tg, of a binary polymer blend in terms of the glass temperatures of the two substituents. The formula is Tg = B1Tg1 + B2Tg2, where Bi is the fraction of flexible bonds of substituent i. A bond is flexible if rotation about it changes the shape of the molecule. Bonds in side groups as well as in the backbone are to be counted. This formula assumes that the free volume, taken here to be the volume fraction of empty lattice sites, is the same for each of the three materials. It has no parameters. The above equation expressed in weight fractions, Wi, is (TgTg1)W111) + (TgTg2)W222) = 0, where ωi is the weight of a monomer unit and ggi is the number of flexible bonds per monomer unit. A more general treatment is given. One variation of the more general treatment which expresses the properties of the blend in purely additive terms gives Tg = B1Tg1 + B2Tg2 + KB1B2(Tg1Tg2)(V01V02), where V0i are the free volume fractions of the homopolymers at their glass temperatures and K is a constant. The added term is usually small. The most general form of the equation requires the energy of interaction between the two unlike molecules, which can be estimated by volume measurements on the blend.  相似文献   

8.
The temperature changes as a result of rapid hydrostatic pressure applications are reported for unvulcanized styrene-butadiene rubber (SBR) in the reference temperature range from 292 to 405 K and in the pressure range from 13.8 to 200 MN m−2. The thermal effects were found to be a function of pressure and temperature. A curve fitting analysis showed that the empirical curve (∂T/∂P)=abP)b−1, described the experimental thermoelastic coefficients obtained from the experiments. The data were analysed by determining the predicted thermoelastic coefficients derived from the Thomson equation (∂T/∂P)=To/Cp. The experimental and the predicted Grüneisen parameter γT were also estimated. Close agreement was found at low pressure but differences were observed at higher pressures between the experimental and expected values for the thermoelastic coefficients and the Grüneisen parameter.  相似文献   

9.
The transported entropy and the Thomson coefficient for charge conducting ions are needed to predict reversible heat effects in batteries. Transported entropies and Thomson coefficients have been calculated from Seebeck coefficients of the cell Fe(s, T1)|Me| β″ alumina | Me | Fe(s, T2) for Na and K (Me). The result is S*Na+ = 56 ± 3 J K−1 mol−1 at 500 K, with a Thomson coefficient τNa+ = 30 ± 2 J K−1 mol−1 in the temperature interval 333–773 K. The transported entropy of Na+ did not change by freezing Na at 370. The results for K+ are identical to those of Na+ within the accuracy of the experiments. The Thomson coefficient derived from measurements at different values of T1 was consistent with the observed variation in emf with ΔT for a given T1. The reversible heat changes at the electrodes have been calculated for sodium sulphur and potassium sulphur batteries. During discharge both batteries produce a net reversible heat, the production always being largest at the alkali metal anode. At the cathode, the heat effect becomes relatively small when the composition of Na and S is within the one phase region. A change in composition from the one phase to the two phase region is expected to lead to changes in local temperature gradients. The systems were described by the electric work method, a method which has practical advantages compared to other electrochemical methods.  相似文献   

10.
The kinetics of the three-phase methanol synthesis over a commercial Cu–Zn–Al2O3 catalyst were studied in an apolar solvent, squalane and a polar solvent, tetraethylene glycol dimethylether (TEGDME). Experimental conditions were varied as follows: P=3.0–5.3 MPa, T=488–533 K and ΦvG/w=7.5×10−3–8×10−3 Nm3 s−1kg−1cat. The nature of the slurry–liquid influences the activation energy and the kinetic rate constant by interaction between adsorbed species and solvent and by competitive adsorption of the solvent on the catalyst surface. The rate of reaction to methanol observed in TEGDME appeared to be about 10 times lower than in squalane. TEGDME reduces the reaction rate, which is a disadvantage for its use as a solvent.  相似文献   

11.
The dielectric relaxation due to absorbed water has been studied with a number of epoxies and other thermosets. In all cases, the absorbed water relaxation strength, as measured by both the dielectric-constant increase and the increase in area under the ″ versus 1/T curve, seems to be attributable to the relaxation of water dipoles and not to Maxwell-Wagner-Silars effects. The activation energies obtained are in the 11–16 kcal mol−1 range. The relaxation strengths observed show that the water molecules manifest roughly 70–100% of their free-state polarizability. In general, the dielectric constant increase per 1 wt% of water absorbed is given to a good approximation by 4.0[(dry) + 2]2 pf/T where f is the fractional polarizability of water in the polymer to its free-state value, found to be 0.7 to 1.0 for most thermosets, and and T are the density and temperature.  相似文献   

12.
Takakazu Kojima  G. C. Berry 《Polymer》1988,29(12):2249-2260
Elastic and quasi-elastic light-scattering, viscometric and rheological studies are given for solutions of the microbial polysaccharide Xanthomonas campestris (xanthan) in aqueous 0.62N NaCl for polymer concentrations from 0.03 to 2.2 g kg−1. The observed negative ∂ln[η]/∂ln T is interpreted as a decrease of the persistence length with increasing T. The behaviour in moderately concentrated solutions (2<[η]c<25) reveals intermolecular association, leading to gel formation in the extreme case. The effect of the association on the viscometric and light-scattering data is discussed. It is concluded that the early stages of association involve structure with the chain axes nearly parallel, but that larger, particulate-like structures develop with increasing c, eventually leading to gel formation under certain conditions.  相似文献   

13.
Surface dynamics during latex film formation   总被引:3,自引:0,他引:3  
Surface dynamics during latex film formation has been investigated theoretically and experimentally by atomic force microscopy. The peak-to-valley distance, y(t), of the latex particles in the surface plane of the latex film decayed exponentially with time during film formation. A theoretical relationship between y(t) and time, t, is given by y(t)=y(0) exp[−t/τ], where y(0) is the value of y(t) when t is zero. τ is a characteristic constant related to the nature of polymer, the particle radius, the surface diffusion coefficient and the temperature. The relationship between the surface diffusion coefficient, Ds, y(0), the radius of the latex particles, R, temperature, T, and τ is given approximately by Ds=1.2×10−20y(0)2[2Ry(0)]2T/τ (cm2/s), where the units are manometers for y(0) and R, kelvin for temperature, and seconds for τ. By measuring the decay of y(t) with time, the surface diffusion coefficient can be obtained. The surface diffusion coefficient for a poly(methyl methacrylate-co-butylacrylate) (50:50) copolymer latex film was found to be A×10−13 cm2/s, A is temperature-dependent.  相似文献   

14.
Results of experiments to determine the distribution of fuel sulphur in one-dimensional pulverized-coal flames are presented for two high volatile A bituminous coals, one pulverized into three size fractions, and one subbituminous A coal. A complete determination of the sulphur fed into the flame was obtained through measurement of SO2, H2S, COS, CS2, and solids sulphur. SO2 is the predominant gas-phase species followed by H2S with COS and CS2 being present in smaller amounts. Sulphur is released from smaller particles more rapidly than from larger ones. Sulphur release parallels, to some extent, nitrogen release except at the longer residence times for the subbituminous coal and one of the bituminous coals. A substantial fraction of the sulphur contained in these coals is inorganic. First-order rate constants for sulphur release from the solid are presented. The effects of internal mass transfer on the kinetics are presented for one of the bituminous coals by considering the sulphur release process to consist of a first-order devolatilization reaction occurring in parallel with diffusion of the sulphur to the particle surface and first-order deposition back to the solid matrix. Particle size effects are accounted for in the model, and the data on solids sulphur are used to obtain activation temperatures for the devolatilization reaction and the deposition reaction of 29 900 and 31 400 K, respectively.  相似文献   

15.
Jer-Yuan Chang  Jin-Long Hong 《Polymer》1998,39(26):7119-7122
Low molecular weight cyanated poly(ether sulfone) (CPES, Mn=3200) was prepared, and cured with bisphenol A dicyanate (BPADCy). The resulting polycyanurates of different compositions show an S-shaped Tg-composition curve. This unexpected S-shaped curve is interpreted in term of the polar interaction between poly(ether sulfone) and s-triazine rings formed by polycyclotrimerization of aromatic dicyanates. Separate study on u.v. spectra of mixture model compounds suggests that this polar interaction belongs to an n–π* interaction between the lone pair electrons of the N-atoms in the s-triazine ring and the π*-orbital of the phenylene rings neighbouring to the sulfone linkage. I.r. study on the cured polycyanurates indicates that this polar interaction causes the shift of the —C=N stretching from 1564 to 1580 cm−1.  相似文献   

16.
For the production of siloxane fluids, the viability of using a multi-channel monolith as a catalyst support system in a three-phase reactor has been studied. The catalyst was tripotassium phosphate (K3PO4). Experiments were performed in a single-channel flow reactor (15 mm i.d. and 500 mm catalyst coated length). The rate of reaction was followed by monitoring the disappearance of the hydroxyl group (–OH). Reaction experiments were performed at a hydroxyl group concentration range from 150 to 170 mol m−3, T=373–413 K and P=7.9 kPa with a nitrogen purge. The maximum temperature of operation was restricted to 413 K to avoid the formation of undesirable by-products. In the regime controlled by chemical kinetics, reaction was of an apparent first order with respect to –OH concentration, and in the apparent rate constant, the pre-exponential factor was 4.19×10−4 ms−1, and the apparent activation energy was 16.1 kJ mol−1. These are only valid for the operating pressure and purge gas flowrate used, as both of these are shown to affect water removal from the liquid phase and, hence, reaction rates. Mass transfer coefficients from the liquid to the catalyst surface were estimated and these increased rapidly with flowrate and were higher than expected for a falling liquid film.  相似文献   

17.
In this paper a series of impedance measurements in the frequency range 10−4−2 × 105 Hz and in the temperature range 20–170°C is reported for the cell: Li-metal/LiSCN [dissolved in poly (ethyleneoxide)]/Li-metal. On the basis of the measurements a whole range of electrical properties such as the conductivity, the charge transfer resistance, the transport number for the Li+-ion, the double layer capacity and the dielectric constant were determined for the polymer complex. The most important findings were (1) two regions with linear log σ vs 1/T plots and a transition temperature between the regions of 80°C and (2) the fact that both ions were mobile in the polymer complex with the Li+-ion having a transport number of 0.54 independent of temperature.  相似文献   

18.
Correlation between the equation of state and the temperature dependence of the self-diffusion coefficient D for polymers such as polystyrene (PS) and polydimethyl siloxane (PDMS) and simple liquids such as argon, methane and benzene and the pressure dependence of D for oligomers such as dimethyl siloxane (DMS) and simple liquids such as cyclohexane and methanol has been examined based on the equation of state derived previously. The experimental data used were published by Antonietti et al. and McCall et al. for polymers, by McCall for linear dimethylsiloxanes and by Jonas et al. and Woolf et al. for simple liquids. The expression for D in this work is given by

where A1(M) is a function of molecular weight Mw, C1(T) and P1(T) are functions of temperature and B1, n1 and m1 are constants determined experimentally. For simple liquids, the values of n1 obtained range from 0.3 to 1.2, with an average , and m1 is in the range 0.5–1.2, with . For polymers, values of n1 are in the range 2.5–7.0 for PS and 0.5–1.3 for PDMS and m1 for DMS is in the range 0.8–1.0. The relation Dη/T = f(M) is found to be useful for simple liquids over a wide range of temperature including the critical region and for pressures up to ≈5 kbar

1 kbar = 100 MPa There is a close correlation between ln(D/T) and p and βT through ln(D/T)ln Dc−1p−β−1T, where Dc is D at the critical temperature and p and βT are the thermal expansion coefficient and compressibility, respectively. The molecular weight dependence of D for polymers and simple liquids is discussed based on the experimental data and recent theory of Doi and Edwards. A new model for the mechanism of self-diffusion in the liquid state is proposed.  相似文献   


19.
E.N. Dalal  K.D. Taylor  P.J. Phillips   《Polymer》1983,24(12):1623-1630
The equilibrium melting temperature T°m of cis-polyisoprene (Hevea natural rubber) has been determined to be 35.5°C at atmospheric pressure. The optical ‘turbidimetric’ technique developed to obtain the melting data, discussed in this paper, utilized unpolarized light and was free of complications (presumably involving melt strain) encountered with polarized light techniques, but was found to be consistent with that and other techniques. The effect of variables such as heating rate and crystallization time were considered. Two independent methods of extrapolation of the Tm data to evaluate T°m produced values in the range T°m = 35.5° ± 0.3°C. The fold surface free energy σe was also estimated, by two independent methods, to be 0.024 J m−2.  相似文献   

20.
A laboratory scale fixed bed coal gasifier was set up to simulate the conditions existing in the devolatilization zone of an air-blown, fixed-bed coal gasifier. Devolatilization behaviour of a subbituminous coal was evaluated in the temperature range 350 °C to 550 °C and at pressures 30, 300 and 375 psig. Three feed coal particle sizes, (−2, +1), (−4, +3) and (−9, +6)mm, were studied. The gas feed was a synthetic mixture of composition similar to that leaving the gasification zone of a fixed bed gasifier and contained 30% by volume of steam. Devolatilization runs were conducted over coal residence times of 5, 10, 20 and 30 min durations. The gas evolution rates showed a peak around 5 min from the start of a run and most of the gas evolution tapered off just under 30 min. Thirty key components in the tars were quantified and these included aliphatic and aromatic homologues, as well as sulphur and nitrogen substituted structures. The molecular weights of the tar samples showed a maximum between 300 and 500. A first order kinetic model applied to the total weight loss data yielded activation energies in the range 4 to 11 kcal mol−1. Differential equations for obtaining concentration profiles for tar and gas inside the coal particle were solved numerically. From these calculations it was concluded that the pressure buildup (due to evolution of tar and gas) inside the coal particle was higher for larger particles, at a given external pressure, but decreased with external pressure. The concentration of tar inside the particle did not appear to be sensitive to low pressures (around 1 atm), but increased in the higher range of pressure (above 20 atm) and also with particle size.  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-23

京公网安备 11010802026262号