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.
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   


2.
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   

3.
Catalytic oxidation of trichloroethylene (TCE) and trichloromethane (TCM) oxidized alone and in two-component mixtures with selected volatile organic compounds (VOCs) such as toluene, n-hexane, ethanol or acetone was investigated over a Pt-Pd-based catalyst on a monolithic, metallic, γ-Al2O3-washcoated support. TCE and TCM were more difficult to oxidize than VOCs and temperatures of their 50% conversion (T50%) amouted to 420 and 330 °C, respectively. All the VOCs added were found to enhance the conversion of the two chlorinated compounds, drecreasing T50% by 20 °C, at the most, for TCM in the presence of toluene and by 50 °C for TCE in the presence of acetone. Both the chlorinated compounds lowered the conversion of the VOCs added (except that of toluene), and this lowering was particularly distinct with n-heptane; they also raised the concentration of acetaldehyde formed during the oxidation of oxyderivative compounds.  相似文献   

4.
Drop size distribution(DSD) or mean droplet size(d32) and liquid holdup are two key parameters in a liquid–liquid extraction process. Understanding and accurately predicting those parameters are of great importance in the optimal design of extraction columns as well as mixer–settlers. In this paper, the method of built-in endoscopic probe combined with pulse laser was adopted to measure the droplet size in liquid–liquid dispersions with a pump-impeller in a rectangular mixer. The dispersion law of droplets with holdup range 1% to 24% in batch process and larger flow ratio range 1/5 to 5/1 in continuous process was studied. Under the batch operation condition, the DSD abided by log-normal distribution. With the increase of impeller speed or decrease of dispersed phase holdup, the d32 decreased. In addition, a prediction model of d32 of kerosene/deionized system was established as d_(32)/D = 0.13(1 + 5.9φ)We~(-0.6). Under the continuous operation condition, the general model for droplet size prediction of kerosene/water system was presented as d_(32)/D = C_3(1 + C_4φ)We~(-0.6). For the surfactant system and extraction system, the prediction models met a general model as d_(32)/D = bφ~nWe~(-0.6).  相似文献   

5.
The rehydration characteristics of potato cylinders were examined. Samples were pre-dried in a convective oven (60 °C) or in a microwave oven (250, 440 or 600 W), and subsequently rehydrated in a water bath at temperatures between 20 and 80 °C. Fick's Second Law of Diffusion was used to describe the rehydration kinetics. The process was characterized by two effective diffusion coefficients (D1, D2), or two stages. The effect of temperature on D was interpreted using the Arrhenius relationship. The rehydration kinetics were dependent on temperature, solid-to-liquid ratio, sample dimensions (diameter and diameter-to-length ratio), pre-blanching, drying method (convective or microwave) and pre-soaking (ionic surfactants and NaCl). Agitation and non-ionic surfactants did not effect the rehydration process.  相似文献   

6.
Mixed oxides of alumina and zirconia having a relative composition of 50, 80 and 100% Zr2O were synthesized by means of sol–gel methods. The catalysts were sulfated with H2SO4 1N, and were loaded with 0.3% Pt metal using the incipient wetness technique. The characterization of the physicochemical properties was carried out using XRD, N2-adsorption at 78 K, and SEM. The catalytic properties of the Al2O3–ZrO2 series were studied by means of dehydration of 2-propanol at 180°C and isomerization of n-hexane at 250°C, 1 atm. The sulfated solids presented a high surface acidity and a limited crystallinity, together with high activity for alcohol dehydration (i.e. 2-propanol). On the other hand, the Al2O3–ZrO2 solid solutions (i.e. those having a 20–80% composition) turned out to be the most active ones for the isomerization of n-hexane.  相似文献   

7.
刘露  任丹妮  刘继三 《化工进展》2019,38(z1):179-185
采用溶剂热方法合成了Co-MOF-74,采用X射线衍射(XRD)、氮气吸附、红外光谱(FTIR)、热重-差热分析(TG-DSC)和扫描电子显微镜(SEM)对其进行表征,研究其吸附正己烷/1-己烯性能。结果表明,优化的Co-MOF-74合成工艺条件为:原料配比n[Co(NO32·6H2O]:n(DHTP):n(THF):n(H2O)=2:1:165:750,晶化温度100℃,晶化时间24h,150℃活化2h。合成Co-MOF-74对1-己烯和正己烷的静态饱和吸附量分别为125.6mg/g和72.9mg/g,BET比表面积为1209m2/g,孔体积为0.41cm3/g,微孔平均孔径为0.66nm。L-F吸附模型能较好地解释Co-MOF-74材料对1-己烯的吸附等温线数据,Co-MOF-74吸附1-己烯的过程是热力学自发放热过程。  相似文献   

8.
It was found that TeCl4 in methylene chloride (containing tetrabutylammonium perchlorate as a supporting electrolyte) is reduced to TeCl2, Te0 and Te−2 in potential applied. Two reduction waves are observed during the reduction of TeCl4 at a platinum rde (Ep1 = 0.08 ± 0.02 V, Ep2 = −1.10 ± 0.02 V) due to the reduction of TeCl4 to Te and Te−2 respectively. A cathodic deposition of tellurium from TeCl4 is followed by anodic stripping wave (Ep1 = 0.42 ± 0.02 V), corresponding to the oxidation of Te to TeCl2. It has been shown that reduction of TeCl4, or TeCl−26 to Te causes coating of the electrode with metallic tellurium, on whose surface chloride ions are strongly bonded. It was found that when the electrolysis solution contains excess of chloride ion with respect to TeIV the peak potentials of the cathodic waves shift to more cathodic values (Ep1 = 0.30 ± 0.02 V, Ep2 = −1.25 ± 0.02 V). After cathodic deposition of at least a monolayer of tellurium from such a solution two anodic waves appear (Ep1 = 0.12 ± 0.03 V, Ep2 = 0.24 ± 0.03 V) which are both due to the oxidation of Te to TeII.  相似文献   

9.
The changes produced in the dielectric permittivity, the transverse piezoelectric coefficient, the k31 coupling factor, the s11E elastic compliance and the mechanical quality factor of PZT-4D hard piezoceramics by heating have been investigated. The ceramics were then repoled, and the reversible and irreversible components of the changes quantified. The results showed that depolarisation began at 150°C. A high level of poling was retained even after heating at 300°C (d31=−83×10−12 C N−1 and k31=0.225), only 20°C below the transition temperature. However, a significant irreversible degradation of the mechanical quality factor, Qm, occurred at a temperature as low as 100°C. Experiments on thinned specimens showed that the degradation of Qm took place in the Ag doped layer produced by the electrodes. Indentation surface cracks were also introduced into the ceramics to investigate the behaviour of cracks during the thermal treatments.  相似文献   

10.
The adsorption dynamics of nitrogen dioxide on activated carbon are measured using a microbalance technique. A theoretical model, incorporating the mechanisms of micropore, macropore and concentration dependent sorbed phase (surface) diffusion, with nonlinear equilibrium (dual Langmuir isotherm), is fitted to the experimental NO2 adsorption dynamics over a range of temperatures, concentrations and particles sizes. The data are well fitted by the model over the temperature range 298-373 K, with typical extracted values of the diffusivities at 298 K being: D/R2μ=0.003 s-11 (micropore), Dp = 0.038 cm2/s (macropore) and Dso=1.0 × 10-6cm2/s (surface) with micropore and surface diffusion activation energies of 14.8 and 11.0 kJ/mole, respectively. The influence of the NO2/N2O4 ratio on the adsorption dynamics is also discussed. At temperature higher than 373 K, significant reaction of NO2 with activated carbon results in consumption of the carbon.  相似文献   

11.
As is well known the catalytic activity of Raney nickel may be increased by alloying with some percentage of transition metals eg Fe, Ti, or Mo. Electrocatalysts activated in this way may reach exchange current densities up to 5 A/g, and in the anodic oxidation of H2 current densities up to 4 A/g at 23°C. Sorption measurements and pore distributions calculated from the sorption isotherms show that the structure of the Raney sponge is changed but little by these transition metal additions; by the observed small structure differences the different properties of electrocatalysts cannot be explained. Discussion of these results shows that the measured high exchange current densities alone do not account for the high anodic current densities, because the diffusion of molecular hydrogen through the transitional pore system of the Raney sponge filled by electrolyte to the reaction zone does not admit current densities of more than 1 A/g. Only additional diffusion of chemisorbed H-Atoms on the catalyst surface allows the high current densities observed. Therefore an essential condition of a good electrocatalyst surface is a large coefficient D0 of the surface diffusion of hydrogen exceeding 5 × 10−9 cm2/s.  相似文献   

12.
13.
The conversion of n-hexane and methanol into value-added aromatic compounds is a promising method for their industrially relevant utilization. In this study, intergrown ZSM-5/ZSM-11 crystals were synthesized and their resulting catalytic performance was investigated and compared to those of the isolated ZSM-5 and ZSM-11 zeolites. The physicochemical properties of ZSM-5/ZSM-11 intergrown zeolite were analyzed using X-ray diffraction, N2 isothermal adsorption-desorption, the temperature-programmed desorption of ammonium, scanning electron microscopy, Fourier transform infrared spectra of adsorbed pyridine, and nuclear magnetic resonance of 27Al , and compared with those of the ZSM-5 and ZSM-11 zeolites. The catalytic performances of the materials were evaluated during the co-feeding reaction of methanol and n-hexane under the fixed bed conditions of 400°C, 0.5 MPa (N2), methanol:꞉n-hexane=7꞉:3 (mass ratio), and weight hourly space velocity=1 h–1 (methanol). Compared to the ZSM-5 and ZSM-11 zeolites, the ZSM-5/ZSM-11 zeolite exhibited the largest specific surface area, a unique crystal structure, moderate acidity, and suitable Brønsted/Lewis acid ratio. The evaluation results showed that ZSM-5/ZSM-11 catalyst exhibited better catalytic reactivity than the ZSM-5 and ZSM-11 catalysts in terms of methanol conversion rate, n-hexane conversion rate, and aromatic selectivity. The outstanding catalytic property of the intergrown ZSM-5/ZSM-11 was attributed to the enhanced diffusion associated with its unique crystal structure. The benefit of using zeolite intergrowth in the co-conversion of methanol and alkanes offers a novel route for future catalyst development.  相似文献   

14.
Pt/Al2O3 catalysts with smaller size of Pt nanoparticles were prepared by ethylene glycol reduction method in two different way and their oxidation activities for three typical VOCs (volatile organic compounds) were evaluated. The catalyst prepared by first adsorption and then reduction procedure is denoted as L-Pt/Al2O3 while the catalyst prepared by first reduction and then loading procedure is defined as R-Pt/Al2O3. The results show that L-Pt/Al2O3 with the stronger interaction between Pt species and Al2O3 exhibit smaller size of Pt nanoparticles and favorable thermal stability compared with R-Pt/Al2O3. L-Pt/Al2O3 is favor of the formation of more adsorbed oxygen species and more Pt2+ species, resulting in high catalytic activity for benzene and ethyl acetate oxidation. However, R-Pt/Al2O3 catalysts with higher proportion of Pt0/Pt2+ and bigger size of Pt particles exhibits higher catalytic activity for n-hexane oxidation. Pt particles in R-Pt/Al2O3 were aggregated much more serious than that in L-Pt/Al2O3 at the same calcination temperature. The Pt particles supported on Al2O3 with~10 nm show the best catalytic activity for n-hexane oxidation.  相似文献   

15.
M. Best  H. Sillescu   《Polymer》1992,33(24):5249-5253
We describe a simple light scattering set-up for measuring interdiffusion coefficients D in polymer blends by generating spinodal decomposition and subsequent dissolution after temperature jumps across the phase boundary. In blends of polystyrene and polymethylstyrene (random copolymer of 60% m-methylstyrene and 40% p-methylstyrene) D values were obtained between 10−11 and 10−15 cm2s−1 at temperatures up to 50 K above the upper critical solution temperature. The results are discussed in relation to tracer diffusion in the same system.  相似文献   

16.
An aerosol dynamics model, AERO2, is presented, which describes the formation of H2SO4-H2O aerosol in a smog chamber. The model is used to analyse how the uncertainties on four input parameters are propagated through an aerosol dynamics model. The input parameters are: the rate of the reaction between SO2 and OH (k1), the ratio between the nucleation rate used in AERO2 and that derived from classical nucleation theory (tn), the H2SO4 mass accommodation coefficient () and a measure of the turbulence intensity in the reactor (ke). Uncertainties for these parameters are taken from the literature. One of the results of the analysis is that AERO2 and aerosol dynamics models in general can only predict upper bounds for the total number (Ntot) and total volume (Vtot) concentrations of the particles. The uncertainties on Ntot and Vtot are mainly due to the uncertainties on k1, and tn. An uncertainty factor of 20–100 still remains when the uncertainty on k1, is reduced to ±5%. Aerosol measurements from three smog chamber experiments have therefore been used, in an attempt to reduce the uncertainty on k1 and tn. Values for k1 are obtained in the reduced range 7.8 × 10−13 to 1.0 × 10−12 cm3 s−1, which is within the range found in the literature. For tn, values in the range 104–107 are obtained, which is close to the upper bound of the range in literature. These values for tn are in marked contrast with a recent set of experiments on nucleation in H2SO4-H2O mixtures, which suggests a value for tn of at most 10−5.  相似文献   

17.
Mass transfer coefficients at cylindrical, H2 evolving electrodes, were measured by determining the reduction rate of K3Fe(CN)6. The variables studied were: gas discharge rate V, diameter of the cylinder D, height and position of the cylinder. The diameters ranged from 0·2–2·5 cm, the cd from 25-380 mA/cm2. For horizontal cylinders, the following correlation was found: log K = a + 2·17 log(V0·11/D0·08). The application of gas evolving cylindrical electrodes in industrial electrolysis is discussed in comparison with rotating electrodes.  相似文献   

18.
S.J. Candau  I. Butler  T.A. King 《Polymer》1983,24(12):1601-1609
The diffusion constant of solutions of polystyrenes of molecular weight ranging from 110 000 to 3.6 × 106 in straight-chain-dibutylphthalate has been measured by photon correlation spectroscopy as a function of polymer volume fraction and temperature. In the semi-dilute range the co-operative diffusion constant Dc exhibits a much smaller increase with the polymer volume fraction Φ than theoretically predicted, except at high temperatures (T 100°C) where it follows a Φ1/2 law characteristic of a marginal solvent. This effect can be described to an enhancement of the friction factor, which is also demonstrated by a decrease in Dc occurring at a volume fraction which increases with temperature. The same effect is observed in swollen networks but it is strongly reduced for swelling equilibrium conditions.  相似文献   

19.
Characteristic electrochemical transport parameters for an experimental poly(ethylene)terephtalate (PET) track-etched membrane with well-defined structure and low porosity (Θ=0.13%) were determined with the membrane in contact with KCl solutions at different concentrations. Membrane potential, Δφm, measurements were performed to investigate the effective fixed charge concentration, Xf, and transport number of the ions, ti, in the membrane using two different procedures: keeping the concentration ratio constant, or keeping one concentration constant and changing the other one. Results show the membrane presents a weak cation-exchanger character, since the following values were obtained: Xf=−(2.5±0.2)×10−2 M, tK =(0.56±0.06), tCl=(0.44±0.05); taking into account these values, concentration dependence of membrane potential was predicted. Membrane electrical resistance, Rm, was obtained from Impedance Spectroscopy (IS) measurements using equivalent circuits as models, and the membrane porosity Θ=(0.11±0.02)% was also obtained from resistance values, which agrees very well with the value determined from geometrical parameters. From Rm, Δφm and Θ values, the diffusion coefficient of the ions in the membrane pores can be calculated, and the following average values were obtained: DK+=(1.9±0.4)×10−9 m2/s and DCl=(0.8±0.2)×10−9 m2/s, but for an average concentration higher than 0.06 M, their values do not differ practically from solution in agreement with the small negative charge previously indicated.  相似文献   

20.
Complete Ni2+ exchange of a single crystal of zeolite X of composition Na92Si100Al92O384 per unit cell was attempted at 73°C with flowing aqueous 0.05 M NiCl2 (pH=4.3 at 23°C). After partial dehydration at 23°C and ≈10−3 Torr for two days, its structure, now of composition Ni2(NiOH)35(Ni4AlO4)2(H3O)46Si101Al91O384 per unit cell, was determined by X-ray diffraction techniques at 23°C (space group Fd , a0=24.788(5) Å). It was refined using all intensities; R1=0.080 for the 236 reflections for which Fo>4σ(Fo), and wR2=0.187 using all 1138 unique reflections measured. At four crystallographic sites, 45 Ni2+ ions were found per unit cell. Thirty of these are at two different site III′ positions. Twenty of those are close to the sides of 12-rings near O–Si–O sequences, where each coordinates octahedrally to two framework oxygens, to three water molecules which hydrogen bond to the zeolite framework, and to an OH ion. The remaining 10 are near O–Al–O sequences; only three members of a likely octahedral coordination sphere could be found. In addition, two Ni2+ ions are at site I, eight are at site I′, and five are at site II. Forty six H3O+ ions per unit cell, 24 at site II′ and 22 at site II, each hydrogen bond triply to six rings of the zeolite framework. Each of the 22 H3O+ ions also hydrogen bonds to a H2O molecule that coordinates to a site III′ Ni2+ ion. Six of the eight sodalite cages each contain four H3O+ ions at site II′; the remaining two each contains a tetrahedral orthoaluminate anion at its center. Each tetrahedral face of each orthoaluminate ion is centered by a site I′ Ni2+ ion to give two Ni4AlO4 clusters. The five site II Ni2+ ions each coordinate to a OH ion. With 46 H3O+ ions per unit cell, the great tendency of hydrated Ni2+ to hydrolyze within zeolite X is demonstrated. With a relatively weak single-crystal diffraction pattern, with dealumination of the zeolite framework, and with an apparent decrease in long-range Si/Al ordering likely due to the formation of antidomains, this crystal like others treated with hydrolyzing cations appears to have been damaged by Ni2+ exchange and partial dehydration.  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