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.
The radiation-induced grafting of styrene vapor to low-density polyethylene film of 0.063 mm thickness was studied at 23°C at a dose rate of 1.98 × 104 rad/hr. The concentration C of monomer in the film was measured as a function of pre-irradiation exposure time to monomer vapor. The concentration-dependent diffusion coefficient of styrene in polyethylene was calculated to be 4.9 × 10?9 exp {2.0C/C0} cm2/sec, where C0 is the saturation concentration of styrene in the film, and a linear boundary diffusion coefficient for styrene vapor into polyethylene film was found to be 2.0 × 10?7 cm/sec. The rate of grafting was determined as a function of the concentration of styrene absorbed in the film. The maximum graft yield was obtained with an initial styrene concentration in the film of 4 wt-%. Under conditions of low initial monomer concentration, the grafting rate increases with irradiation time. The results are compared with previously published data on grafting of polyethylene from methanol–styrene solutions. They are explained in terms of the viscosity of the amorphous region as a function of styrene content and the resistance to the diffusion of monomer at the film–vapor interface.  相似文献   

2.
Drawing of linear polyethylene at 60°C. to an extension ratio of ten drastically reduces the sorption and diffusion of n-pentane, benzene, methylene chloride, and tetrachloroethylene. Methylene chloride was chosen for more detailed study. The sorption is of the normal Fickean type. It is also fully reversible in the temperature range between 25 and 45°C. if the sorbed amount is kept to below 0.5%. At higher concentrations the sample relaxes so that sorption irreversibly increases. The reversible sorption per gram of amorphous component is about 1/6 of that in undrawn polyethylene. The diffusion constant has a larger temperature and concentration dependence than in the undrawn material. At zero concentration the activation energy for diffusion is 34.4 kcal./mole and the diffusion constant at 25°C. is 8 × 10?11 cm.2/sec. as compared with 14.4 kcal./mole and 1.5 × 10?8 cm.2/sec. in undrawn PE. Cold drawing reduces the sorption sites without changing their energy content, but drastically cuts down diffusion and increases the activation energy. A smaller part of the increase of the latter is a consequence of the lower enthalpy of the amorphous material and a larger part is probably due to the increased distance between sorption sites.  相似文献   

3.
The fluorescence quenching of polystyrene by oxygen at room temperature was investigated. The diffusion coefficient of oxygen in polystyrene films was taken as 3.1 × 10?7 cm2/sec; the energy migration coefficient of polystyrene in the excited singlet state was estimated as 6 × 10?6 cm2/sec.  相似文献   

4.
A comprehensive evaluation of Cs ions sorption to and diffusion in crushed granite was conducted in this study. The sorption capacity of crushed granite suggested by the Langmuir model was 5.48 × 10?6 mol‐Cs/g‐granite. The distribution coefficient (Kd) was around 7.5 mL/g and pH independent. By using an in‐diffusion method with a modified capillary column, some diffusion relevant parameters of Cs ions in crushed granite were derived. The apparent diffusion coefficient (Da) seemed unaffected by Cs concentration (1.15 × 10?10 to 2.82 × 10?10 m2/s, at 10?7 and 10?3 M, respectively). The determined effective diffusion coefficients (De) were located in the window from 8.59 × 10?10 (10?7 M) to 1.69 × 10?9 (10?3 M) m2/s. Under various pH environments, pH independent Da (9.0 × 10?9 m2/s) and De (1.0 × 10?9 m2/s) values were observed. Under current systems, consistently higher De than Da implied the diffusion of Cs ions was governed by surface diffusion phenomenon. Whereas the pH insensitive feature indicated the Cs sorption to crushed granite was mainly through ion‐exchange reaction. Moreover, further SEM/EDS mapping clearly showed the adsorbed Cs ions were highly concentrated on the fracture surface of biotite.  相似文献   

5.
钴离子在P507为载体的支撑液膜中的传输   总被引:5,自引:0,他引:5  
本文用国内的膜材料设计了以2-乙基已基膦酸单(2-乙基已基)酯(HEH(EHF),P_(507))为载体的支撑液膜迁移钴的实验,获得的数据为深入实验提供了依据.基本原理支撑液膜(SLM)是依靠微孔塑料薄膜(支撑体)微孔的毛细管作用将含萃取剂(载体)的有机溶液吸附在支撑体上.P_(507)萃取钴表示为:  相似文献   

6.
The onset of micelles formations critical micelle concentration, diffusion coefficients as well as particle sizes for some new synthesized anionic copolymer surfactants PSA4a, PSA4b, and PSA4c, were determined and discussed. Three different electrochemical techniques such as cyclic voltammetry (CV), rotating disk voltammetry (RDV), and chronocoulometry (CC) were used in this investigation. The voltammetry of electroactive hydrophobic probe ferrocene solubilized surfactants was investigated in aqueous buffer carbonate solutions of pH 10. The CMC for each PSA4a, PSA4b, and PSA4c, was found to be 3.20 × 10?4, 4.60 × 10?4 and 6.30 × 10?4 M, respectively, using both CV and RDV techniques. The amount of adsorption contribution of ferrocene solubilized surfactants at the glassy carbon electrode was determined from chronocoulometric measurements and it was found in the range from (1.4 to 2.7) × 10?15 M. The apparent diffusion coefficients were estimated from RDV measurements and the real micelles diffusion coefficients were obtained. Re-quilibrium considerations of ferrocene probe kinetics at the electrode surface were treated according to two different modes of slow- and fast-kinetics. The corrected diffusion coefficient values showed constancy at (5.3 ± 0.1) × 10?7, (4.8 ± 0.1) × 10?7, and (3.6 ± 0.4) × 10?7 cm2/sec for PSA4a, PSA4b, and PSA4c, respectively in the concentration range from 20 to 200 mM. The morphological features of anionic copolymeric surfactants PSA4a, PSA4b, and PSA4c, micelles showed globular self-assembled structure.  相似文献   

7.
The vertical concentration profile of 137Cs in cores of marine sediment of Mumbai Harbor Bay has been studied by the compartment and the diffusion-convection models. Based on the measured concentrations of 137Cs in the sediment layers, the various transport parameters such as sedimentation rate, residence half- time, effective migration velocity, diffusion coefficient and the convective velocities were determined. The sedimentation rate was determined to be 1.61, 1.03 0.69 and 1.25 cm year?1 from the slope of lines obtained from the depth profile of 137Cs in cores using a least-square fitting method at site 1, 2, 3 and 4 respectively. The mean residence half-times, ranging from 11 to 35 years were observed to be the highest at the upper layers (up to 8 cm) of all sites and decreased with sediment depth. Subsequently, the ranges of mean value of effective vertical migration velocity in the same layers were between 0.15 and 0.46 cm year?1. As expected, the vertical migration in the upper sediment layers was very slow and thereafter increased slowly in the succeeding layers (12 cm onwards) of all sites with a mean ranging from 1.11 to 4.13 cm year?1. The obtained migration velocities were quite higher than those reported in literatures for global fallout. The convective velocity and diffusion coefficient at each site were assumed to be constant in the whole depth and calculated under the assumption of steady state. Using a depth-zoned bioturbational mixing model, the estimated biological diffusion coefficients ranged from 7 × 10?7 to 3.8 × 10?6 cms?1 which were within the literature values reported for shallow coastal environments and deep sea.  相似文献   

8.
Values of the translational diffusion coefficient of a spin probe (2,2,6,6-tetramethyl-4-oxypiperidine-1-oxyl) are determined via cyclic voltammetry, chronoammetry, and EPR in ionic liquid (1-butyl-3methylimidazolium tetrafluoroborate) at 295 K. They are found to be (8 ± 3) × 10?12, (9 ± 3) × 10?12, and (11 ± 4) × 10?12 m2/s, respectively.  相似文献   

9.
Diffusion coefficients and solubilities of methane in polyisobutylene have been measured at four temperatures between 102 and 188°C. in the pressure range 23–341 atm. Diffusion coefficients extrapolated to atmospheric pressure range from 1.72 × 10?6 cm.2/sec. at 102°C. to 1.5 × 10?5 cm.2/sec. at 188°C. corresponding to an activation energy for diffusion of 8.7 ± 0.4 kcal./mole. Solubilities are small, about one molecule of methane for every forty carbon atoms in the polyisobutylene at 300 atm. partial pressure of methane. Solubilities vary little with temperature, but show an apparent minimum between 127 and 188°C. With improved methods of data analysis, diffusion coefficients and solubilities have been recalculated from previously reported studies on nitrogen in branched polyethylene and methane in branched polyethylene, linear polyethylene, and polystyrene. Recalculated diffusion coefficients are essentially the same as those reported previously, but the recalculated solubilities are decreased from 2 to 30%. The solubilities of all five systems show strong deviations from Henry's law, i.e., increases in partial pressure of methane and nitrogen with respect to solubility exceed linearity. The partial pressure (or fugacity)—solubility data may be interpreted in terms of a sorption model in which sorbed molecules are accommodated in widely dispersed, unoccupied volumes or sites in the polymer. An almost equivalent, solution model in which the first sorbed molecules to enter the polymer are accommodated to a large extent in existing volumes in the polymer, with successively sorbed molecules swelling the polymer to a greater extent (i.e., partial molal volume of sorbed molecules, V 1, increasing with concentration) can also account for these data.  相似文献   

10.
Diffusion coefficients of ethylene glycol (EG) have been measured in poly(ethylene terephthlate) (PET) melts by a quartz-spring sorption apparatus. A simple mathematical model was developed to investigate the sorption behavior accompanied by chemical reactions of EG and PET at high temperatures. Diffusion coefficients are deduced from experimental data for an asymptotically thin sample in order to minimize the effects of reactions. The diffusion coefficient of EG is strongly dependent on the vapor pressure of EG and temperature but not on the molecular weight of PET in this experimental range (degree of polymerization 80–120). The diffusion coefficient of EG in PET melt at 265°C is 2.58 × 10?7 cm2/s at the limit of zero concentration of EG. The activation energy for diffusion is 38.4 kcal/gmol, and the heat of solution for sorption is ?44.9 kcal/gmol. The concentrations of the volatile materials resulting from reactions in PET-EG system were analyzed with gas chromatography. In addition, a fit of the current model to experimental data yields frequency factors for the polymerization reaction (k1) and the acetaldehyde formation reaction (k2) to be 5.84 × 108 cm3/mol ? min and 3.90 × 1011 min?1, respectively.  相似文献   

11.
This study has investigated the migration of stabilizers from three polypropylene materials: a polypropylene homopolymer, a polypropylene block copolymer, and a polypropylene random copolymer in 10% ethanol, 3% acetic acid, 20% ethanol, 50% ethanol, and isooctane according to Regulation (EU) No. 10/2011. Measurements were performed at 20, 40, and 70°C and migration was evaluated from 10?min to 10 days. Measurements were performed by high-performance liquid chromatography (HPLC) with external calibration. The HPLC method provided high correlation coefficients, good precision, good accuracy, and suitable reproducibility. Diffusion coefficients for stabilizers were obtained using a rigorous model based on Fick’s second law and the values were between 6.1?×?10?13 and 3.8?×?10?9?cm2?s?1. By applying an Arrhenius-type equation to the diffusion coefficients, an estimation of activation energy of the diffusion was obtained. The activation energies were from 39.97 to 98.75?kJ?mol?1 for the stabilizers. The results indicate that the polypropylene material structure influenced the migration rate, which decreased in the order of the increasing crystallinity in the materials. The diffusion coefficients for stabilizers in the polypropylene random copolymer were higher than in the polypropylene block copolymer and polypropylene homopolymer. The random polypropylene copolymer had the lowest activation energy. These results show that a higher diffusion coefficient indicates a higher migration rate, and a system with a lower activation energy allows more migration.  相似文献   

12.
The initiation of the polymerization of acrylamide by 4-4′-dicyano-4-4′-azopentanoic acid in aqueous solution has been studied kinetically at 25°C. Ferric chloride and ferric sulfate were used to terminate polymerization so that rates of initiation could be calculated from the rates of production of ferrous iron. Velocity coefficients at 25°C. for the initiation reaction were found to be (25.7 ± 2.8) × 10?7 sec.?1 for the ferric chloride terminated reaction and (73.6 ± 0.6) × 10?7 sec.?1 for the ferric sulfate-terminated polymerization. The value reported for the initiation reaction when acrylamide is polymerized in the absence of metal salts is 1.29 × 10?7 sec.?1. Velocity coefficients for the termination reaction have been calculated from the overall rates of polymerization obtained with ferric salts present. In the case of the ferric chloride-terminated reaction, it has been shown that the rate of polymerization is reduced by increasing the total concentration of chloride ions. Termination velocity coefficients at 25°C. for the inner sphere complexes FeCl2+·5H2O and FeSO4+·4H2O have been calculated to be 18.9 × 104 and 7.98 × 104 l./mole-sec., respectively. The dependence on the concentration of ferric chloride of the molecular weights of the polymers produced has also been considered.  相似文献   

13.
The sorption of bromine from bromine water on cis-1,4-polyisoprene film and the initial stages of bromination are studied at concentrations of 0.002–0.1 mol/l and temperatures of 15–35°C. The diffusion coefficient of bromine into natural rubber (NR) is 1.3–2.0 × 10?6 cm2/sec for the total sorption and 5–13 × 10?7 cm2/sec for the irreversible sorption. The partition coefficient of bromine between water and rubber increased from 17.3 at 15°C to 37.1 l/kg at 35°C. The chemical potential, enthalpy and change in entropy of partition are, at 25°C, respectively: ?1.9 kcal/mol, 6.6 kcal/mol and 28.4 cal/mol. K. The irreversible sorption is due to a charge-transfer complex between bromine molecules and double bonds of the rubber. The complex is the first stage of the addition reaction, which becomes noticeable at concentrations above 0.012 mol/l. With increasing bromine concentration the concentration of the complex decreases and the added bromine increases. The charge transfer complex appears to change the conformation of the cis-NR chains so that the bromine addition occurs in the trans-conformation, as shown by FT–IR spectra. The bromination is accompanied by a marked crystallization effect as illustrated by thermal analysis and WAXS measurements.  相似文献   

14.
The electrocatalytic oxidation of hydrazine at a carbon paste electrode spiked with acetylferrocene as a mediator was studied by cyclic voltammetry, differential pulse voltammetry, and chronoamperometry. In contrast to other ferrocenic compounds, acetylferrocene exhibits a chemical irreversible behavior, but it can act as an effective mediator for electrocatalytic oxidation of hydrazine, too. The heterogeneous electron transfer rate constant between acetylferrocene and the electrode substrate (carbon paste) and the diffusion coefficient of spiked acetylferrocene in silicon oil were estimated to be about 3.45×10?4 cm s?1 and 4.45×10?9 cm2 s?1, respectively. It has been found that under the optimum conditions (pH 7.5) the oxidation of hydrazine occurs at a potential of about 228 mV less positive than that of an unmodified carbon paste electrode. The catalytic oxidation peak current of hydrazine was linearly dependent on its concentration and the obtained linear range was 3.09×10?5 M–1.03×10?3 M. The detection limit (2σ) has been determined as 2.7×10?5 M by cyclic voltammetry. Also, the peak current was increased linearly with the concentration of hydrazine in the range of 1×10?5 M–1×10?3 M by differential pulse voltammetry with a detection limit of 1×10?5 M. This catalytic oxidation of hydrazine has been applied as a selective, simple, and precise new method for the determination of hydrazine in water samples.  相似文献   

15.
《Electroanalysis》2006,18(17):1703-1709
A simple bipartite expression for transient amperometric currents at a membrane covered planar electrode has been derived. This expression is uniquely applicable across the entire measurement time frame and enables a rapid, comprehensive dynamic simulation of solute diffusion through the membrane barrier. The effective diffusion coefficient through the membrane for an electrochemically active solute can be readily obtained by employing the simple curve fitting routine. The diffusion coefficients for 4‐acetaminophenol through 0.05 and 0.025 μm pore diameter mixed cellulose esters membranes were determined to be 1.96×10?7 and 1.21×10?7 cm2/s, respectively. For catechol respective values were 1.55×10?7 and 1.15×10?7 cm2/s. Diffusion coefficients were reduced following membrane exposure to bovine serum albumin and this change provided a functional counterpart to conventional structural measures of surface biofouling. The approach has relevance to both sensors and membrane separation systems in biomedicine, and has the potential to offer a rapid assessment of membrane surface biofouling.  相似文献   

16.
The Henry's law and diffusion constants of vinyl chloride in poly(vinyl chloride) were determined at temperatures of 24, 90, 120, 150, and 170°C for weight fractions of vinyl chloride between 0.2 × 10?3 and 0.8 × 10?3. Above 90°C, Henry's law applies; values of the constant increase with temperature from 1.8 × 102 to 5.5 × 102 atm per unit weight fraction of dissolved vinyl chloride. The heat of desorption is about 15 kJ/mole. At 24°C, the nominal Henry's law constant was smaller than would have been obtained by extrapolating the values found at higher temperature. The diffusion constants increase with temperature from about 2 × 10?13 to 3 × 10?7 cm2/sec. The activation energy for diffusion is about 110 kJ/mole between 90 and 170°C. Although all values were determined in the absence of air, it is likely that they apply to polymer in air. They may, therefore, be used to calculate the vinyl chloride content in the gas above poly(vinyl chloride) under specific processing conditions.  相似文献   

17.
Based on single‐walled carbon nanotubes (SWCNTs) modified glassy carbon electrode (GCE/SWCNTs), a novel method was presented for the determination of L ‐tyrosine. The GCE/SWCNTs exhibited remarkable catalytic and enhanced effects on the oxidation of L ‐tyrosine. In 0.10 mol/L citric acid‐sodium citrate buffer solution, the oxidation potential of L ‐tyrosine shifted negatively from +1.23 V at bare GCE to +0.76 V at GCE/SWCNTs. Under the optimized experimental conditions, the linear range of the modified electrode to the concentration of L ‐tyrosine was 5.0×10?6–2.0×10?5 mol/L (R1=0.9952) and 2.7×10?5–2.6×10?4 mol/L (R2=0.9998) with a detection limit of 9.3×10?8 mol/L. The kinetic parameters such as α (charge transfer coefficient) and D (diffusion coefficient) were evaluated to be 0.66, 9.82×10?5 cm2 s?1, respectively. And the electrochemical mechanism of L ‐tyrosine was also discussed.  相似文献   

18.
Nonlinear optical absorption at a wavelength of 1080 nm for nanocomposite thin films made from polymers and single-walled carbon nanotubes (SWNTs) was studied by the longitudinal scanning (z-scan) technique. Two SWNT types differing in the synthesis procedure (HipCO and arc evaporation techniques) were used for the preparation of nanocomposites based on the polymers carboxymethylcellulose (CMC) and poly(vinyl alcohol) (PVAL). The nonlinear absorption coefficients were measured to be ?5.0 × 10?7 and ?3.9 × 10?7 cm/W for the CMC-SWNT/HipCO and CMC-SWNT/Arc composite films and ?1.7 × 10?7 and ?0.9 × 10?7 cm/W for the PVAL-SWNT/HipCO and PVAL-SWNT/Arc films, respectively. It was found that the film nanocomposites based on carboxymethylcellulose had a higher absolute value of the nonlinear absorption coefficient than the films in which PVAL was used as the polymer matrix.  相似文献   

19.
《Electroanalysis》2006,18(12):1193-1201
A chemically modified carbon paste electrode with 2,7‐bis(ferrocenyl ethyl)fluoren‐9‐one (2,7‐BFEFMCPE) was employed to study the electrocatalytic oxidation of ascorbic acid in aqueous solution using cyclic voltammetry, differential pulse voltammetry and chronoamperometry. The diffusion coefficient (D=1.89×10?5 cm2 s?1), and the kinetic parameter such as the electron transfer coefficient, α (=0.42) of ascorbic acid oxidation at the surface of 2,7‐BFEFMCPE was determined using electrochemical approaches. It has been found that under an optimum condition (pH 7.00), the oxidation of ascorbic acid at the surface of such an electrode occurs at a potential about 300 mV less positive than that of an unmodified carbon paste electrode. The catalytic oxidation peak currents show a linear dependence on the ascorbic acid concentration and linear analytical curves were obtained in the ranges of 8.0×10?5 M–2.0×10?3 M and 3.1×10?5 M–3.3×10?3 M of ascorbic acid with correlation coefficients of 0.9980 and 0.9976 in cyclic voltammetry and differential pulse voltammetry, respectively. The detection limits (2δ) were determined to be 2.9×10?5 M and 9.0×10?6 M with cyclic voltammetry and differential pulse voltammetry, respectively. This method was also examined for determination of ascorbic acid in pharmaceutical preparations.  相似文献   

20.
The electrochemical behavior of Yb3+ in perchloric media was studied by cyclic voltammetry and current reversal chronopotentiometry at several temperatures. The results show that the reversible electrochemical reduction of Yb3+ is followed by homogeneous reactions. The experimentally determined diffusion coefficients of ytterbic ion are 0.41×10?5, 0.48×10?5 and 0.53×10?5 cm2 s?1 at 4.2, 9.2 and 14.7°C, respectively, in 0.5 M NaClO4 solutions. From these data a value of 16 kJ mol?1 (3830 cal mol?1) was obtained for the activation energy for diffusion of ytterbic ion. From the activation energy the diffusion coefficient of ytterbic ion at 25.0°C was estimated. The value of 0.67×10?5 cm2 s?1 was obtained. In all the experiments the initial pH was maintained at 4.1.  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