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 sorption—desorption kinetics of nicotinic and isonicotinic acids on strong acid cationites (sulfocationite CU-2-8, sulfonated polycalixarene) were considered. The rate of the processes, following both the cation exchange mechanism H+ ? Cat+ and the ion exchange mechanism accompanied by the chemical reaction of protonation of pyridinecarboxylic acid zwitterions, is determined by a slow diffusion of components in the polymeric phase. The interdiffusion coefficient in sulfocationite CU-2-8 for the process H+ ? Cat+ was equal to ~1?10–11 m2 s–1, in sulfonated polycalixarene the corresponding value was in the range of (1—4)?10–12 m2 s–1.  相似文献   

2.
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   

3.
The photocurrent kinetics in acid solutions have been investigated. The diffusion coefficients of atoms H?((7±2)×10?5cm2s?1) and D?((4±1)×10?5cm2s?1) and OH? and OD? radicals ((1±0.3)×10?5cm2s?1) are found. The rate constants of capture of solvated electrons by H3O+ and D3O+ ions are identical and equal to (8±1)×109M?1s?1. From the shape of the kinetic curves it follows that electrochemical desorption of atomic hydrogen occurs from the adsorbed state. The rate constant of this process has been measured. It is shown that the rate constant of electrochemical desorption depends only slightly on the potential.  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.
The proton magnetic resonance spectrum of lanthanum nickel hydride LaNi5.3H6 was measured over the temperature range 118°K < T < 300°K. The second moment of the absorption at 118°K is M2 = 13.4 ± 0.3 G2. Several possible arrangements of the hydrogen atoms are discussed. Narrowing of the line above 140°K is analyzed in terms of proton diffusion and gives an activation enthalpy E = 21 ± 1 kJ mol?1, NMR correlation time pre-exponential 0.2 ps < τc0 < 1.6 ps and a self diffusion coefficient at 300°K of 2 × 10?12 m2 s?1 < D < 2 × 10?11 m2 s?1.  相似文献   

6.
Interfacial tension (IFT) (γift, N m?1) of benzene‐water; and surface tensions (γ, N m?1) and viscosities (η, N s m?2) of solvents methanol, ethanol, glycerol, ethyl acetate, n‐hexane, diethyl ether, chloroform, benzene, carbon tetrachloride [CCl4], formic acid, Acetonitril, and dimethylformamide [DMF] were measured with Survismeter‐IFT. The ± 1.1 × 10?5 N m?1, ± 1.3 × 10?5 N m?1, and ± 1.1 × 10?5 N s m?2 deviations in respective values were noted. It has 10 times better accuracy than those of individual methods. The survismeter is inexpensive minimizing 2/3 each of consumables, human efforts, time, and infrastructure, cutting down 80% of the waste disposed the environment. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

7.
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   

8.
The self‐diffusion (Dc) coefficients of various lanthanum(III) diamagnetic analogues of open‐chain and macrocyclic complexes of gadolinium used as MRI contrast agents were determined in dilute aqueous solutions (3–31 mM ) by pulsed‐field‐gradient (PFG) high‐resolution 1H‐NMR spectroscopy. The self‐diffusion coefficient of H2O (Dw) was obtained for the same samples to derive the relative diffusion constant, a parameter involved in the outersphere paramagnetic‐relaxation mechanism. The results agree with an averaged relative diffusion constant of 2.5 (±0.1)×10?9 and of 3.3 (±0.1)×10?9 m2 s?1 at 25 and 37°, respectively, for 'small' contrast agents (Mr 500–750 g/mol), and with the value of bulk H2O (2.2×10?9 and 2.9×10?9 m2 s?1 at 25° and at 37°, respectively) for larger complexes. The use of the measured values of Dc for the theoretical fitting of proton NMRD curves of gadolinium complexes shows that the rotational correlation times (τR) are very close to those already reported. However, differences in the electronic relaxation time (τSO) at very low field and in the correlation time (τV) related to electronic relaxation were found.  相似文献   

9.
Tanja Djekic 《Adsorption》2007,13(3-4):231-237
The scope of this work is to determine the effective intraparticle diffusion coefficient of CoCl2 over mesoporous functionalized silica. Silica is selected as a carrier of the functionalized groups for its rigid structure which excludes troublesome swelling, often found in polymeric adsorbents. 2-(2-pyridyl)ethyl-functionalized silica is selected as a promising affinity adsorbent for the reversible adsorption of CoCl2. The adsorption kinetics is investigated with the Zero Length Column (ZLC) method. Initially, experiments were performed at different flow rates to eliminate the effect of external mass transfer. The effect of pore size (60 Å and 90 Å), particle size (40?10?6 m–1000?10?6 m) and initial CoCl2 concentration (1 mol/m3–2.0 mol/m3) on the mass transfer was investigated. A model was developed to determine the pore diffusion coefficient of CoCl2 by fitting the experimental data to the model. The pore diffusion coefficients determined for two different pore sizes of silica are D p (60 Å) =1.95?10?10 [m2/s] and D p (90 Å) =5.8?10?10 [m2/s]. The particle size and the initial CoCl2 concentration do not have an influence on the value of diffusion coefficient. However, particle size has an influence on the diffusion time constant. In comparison with polymer adsorbents, silica based adsorbents have higher values of diffusion coefficients, as well as a more uniform and stable pore structure.  相似文献   

10.
The temperature dependence of the rate coefficients for the OH radical reactions with toluene, benzene, o-cresol, m-cresol, p-cresol, phenol, and benzaldehyde were measured by the competitive technique under simulated atmospheric conditions over the temperature range 258–373 K. The relative rate coefficients obtained were placed on an absolute basis using evaluated rate coefficients for the corresponding reference compounds. Based on the rate coefficient k(OH + 2,3-dimethylbutane) = 6.2 × 10?12 cm3 molecule?1s?1, independent of temperature, the rate coefficient for toluene kOH = 0.79 × 10?12 exp[(614 ± 114)/T] cm3 molecule?1 s?1 over the temperature range 284–363 K was determined. The following rate coefficients in units of cm3 molecule?1 s?1 were determined relative to the rate coefficient k(OH + 1,3-butadiene) = 1.48 × 10?11 exp(448/T) cm3 molecule?1 s?1: o-cresol; kOH = 9.8 × 10?13 exp[(1166 ± 248)/T]; 301–373 K; p-cresol; kOH = 2.21 × 10?12 exp[(943 ± 449)/T]; 301–373 K; and phenol, kOH = 3.7 × 10?13 exp[(1267 ± 233)/T]; 301–373 K. The rate coefficient for benzaldehyde kOH = 5.32 × 10?12 exp[(243 ± 85)/T], 294–343 K was determined relative to the rate coefficient k(OH + diethyl ether) = 7.3 × 10?12 exp(158/T) cm3 molecule?1 s?1. The data have been compared to the available literature data and where possible evaluated rate coefficients have been deduced or updated. Using the evaluated rate coefficient k(OH + toluene) = 1.59 × 10?12 exp[(396 ± 105)/T] cm3 molecule?1 s?1, 213–363 K, the following rate coefficient for benzene has been determined kOH = 2.58 × 10?12 exp[(?231 ± 84)/T] cm3 molecule?1 s?1 over the temperature range 274–363 K and the rate coefficent for m-cresol, kOH = 5.17 × 10?12 exp[(686 ± 231)/T] cm3 molecule?1 s?1, 299–373 K was determined relative to the evaluated rate coefficient k(OH + o-cresol) = 2.1 × 10?12 exp[(881 ± 356)/T] cm3 molecule?1 s?1. The tropospheric lifetimes of the aromatic compounds studied were calculated relative to that for 1,1,1-triclorethane = 6.3 years at 277 K. The lifetimes range from 6 h for m-cresol to 15.5 days for benzene. © 1995 John Wiley & Sons, Inc.  相似文献   

11.
The kinetics of sorption on zirconium hydrophosphates with various water contents was studied. The self-diffusion coefficients of Ni2+ and H+ and the effective diffusion coefficients of Ni2+ ions corresponding to the exchange Ni2+ → H+ were determined. The self-diffusion coefficients of ions in samples with an 85% water content were as high as 4.17 × 10?11 m2/s (Ni2+) and 5.06 × 10?10 m2/s (H+). Water loss caused by drying the sorbents resulted in a decrease in the rate of ion exchange.  相似文献   

12.
A method of measuring the kinetics of currents arising at the electron photoemission from a metal into electrolyte solution when affected by the u.v. laser pulses for 10?8 s at the frequency of repetitions 10–25 Hz is described. Measurements have been taken in solutions without acceptors and in those containing N2O and NO2?, NO3? ions as electron acceptors. The rate constants of capture of the solvated electrons by N2O ((6±1)×09 mol?1 s?1) and NO2? ((4.5±1)×109 mol?1 s?1) and the diffusion coefficients of OH-radicals ((1.0±0.3)×10?5 cm2 s?1) and of NO ((1.2±0.3)×10?5 cm2 s?1) are found. The oxidation rate of NO32? has been shown to decrease from 40 cm s?1 in the range of potentials ?0.55 to ?1.0 V. The rate constant of bimolecular recombination of the solvated electrons ((1.3±0.4)×1010 mol?1 s?1) has been found from the dependence of the emitted charge on the light intensity.  相似文献   

13.
Rate constants have been determined at 296 ± 2 K for the gas phase reaction of NO3 radicals with a series of aromatics using a relative rate technique. The rate constants obtained (in cm3 molecule?1 s?1 units) were: benzene, <2.3 × 10?17; toluene, (1.8 ± 1.0) × 10?17; o? xylene, (1.1 ± 0.5) × 10?16; m? xylene, (7.1 ± 3.4) × 10?17; p? xylene, (1.4 ± 0.6) × 10?16; 1,2,3-trimethylbenzene, (5,6 ± 2.6) × 10?16; 1,2,4-trimethylbenzene (5.4 - 2.5) × 10?16; 1,3,5-trimethylbenzene, (2.4 ± 1.1) × 10?16; phenol, (2.1 ± 0.5) × 10?12; methoxybenzene, (5.0 ± 2.8) × 10?17; o-cresol, (1.20 ± 0.34) × 10?11; m-cresol, (9.2 ± 2.4) × 10?12; p-cresol, (1.27 ± 0.36) × 10?11; and benzaldehyde, (1.13 ± 0.25) × 10?15. These kinetic data, together with, in the case of phenol, product data, suggest that these reactions proceed via H-atom abstraction from the substituent groups. The magnitude of the rate constants for the hydroxy-substituted aromatics indicates that the nighttime reaction of NO3 radicals with these aromatics can be an important loss process for both NO3 radicals and these organics, as well as being a possible source of nitric acid, a key component of acid deposition.  相似文献   

14.
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   

15.
Heterogeneous reactions of gaseous methanesulfonic acid (MSA) with calcium carbonate (CaCO3) and kaolinite particles at room temperature were investigated using diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and ion chromatography (IC). Methanesulfonate (MS?) was identified as the product in the condensed phase, in accordance with the product of the reaction of gaseous MSA with NaCl and sea salt particles. When the concentration of gaseous MSA was 1.34 × 1013 molecules cm?3, the uptake coefficient was (1.21 ± 0.06) × 10?8 (1σ) for the reaction of gaseous MSA with CaCO3 and (4.10 ± 0.65) × 10?10 (1σ) for the reaction with kaolinite. Both uptake coefficients were significantly smaller than those of the reactions of gaseous MSA with NaCl and sea salt particles.  相似文献   

16.
刘佩芳  文利柏 《中国化学》1998,16(3):234-242
The mass transport and charge transfer kinetics of ozone reduction at Nafion coated Au electrodes were studied in 0.5 mol/L H2SO4 and highly resistive solutions such as distilled water and tap water. The diffusion coefficient and partition coefficient of ozone in Nafion coating are 1.78×10-6 cm2·s-1 and 2.75 at 25℃ (based on dry state thickness), respectively. The heterogeneous rate constants and Tafel slopes for ozone reduction at bare Au are 4.1×10-6 cm·s-1, 1.0×10-6 cm·s-1 and 181 mV, 207 mV in 0.5 mol/L H2SO4 and distilled water respectively and the corresponding values for Nafion coated Au are 5.5×10-6 cm·s-1, 1.1×10-6 cm·s-1 and 182 mV, 168 mV respectively. The Au microelectrode with 3 μm Nafion coating shows good linearity over the range 0-10 mmol/L ozone in distilled water with sensitivity 61 μA·ppm-1 ·cm-2, detection limit 10 ppb and 95% response time below 5 s at 25℃. The temperature coefficient in range of 11-30℃ is 1.3%.  相似文献   

17.
The electrooxidation of dextromethorphan on a composite constructed with carbon nanotube–ionic liquid–carbon microparticles was investigated by cyclic voltammetry in a 100 mM phosphate buffer solution, pH 7.40. In the voltammograms, an irreversible diffusion-controlled anodic peak appeared. The diffusion coefficient of dextromethorphan, the electron-transfer coefficient, and the standard rate constant of the electrooxidation process were found to be 3.45?×?10?6 cm2 s?1, 0.65, and 1.67?×?10?3 cm s?1, respectively. A sensitive and timesaving determination procedure was developed for the analysis of dextromethorphan, and the corresponding analytical parameters were reported. Using this method, dextromethorphan was determined with an LOD and LOQ of 8.81 and 29.36 μM in a linear range of 2.5?×?10?4 to 3.3?×?10?3 M, respectively. The proposed amperometric method was successfully applied to the analysis of commercial pharmaceutical products (syrup and oral drop), and the results were in good agreement with the declared values.  相似文献   

18.
This study investigates sorption and diffusion of Cs in two potential host rocks (granite from Kinmen Island and basalt from Penghu Island) by using batch and through-diffusion methods in order to establish a reliable safety assessment methodology. These methods were applied to crushed and intact rock samples to investigate the actual geological environment. According to solid-phase analysis, including X-ray diffraction, elemental analysis, auto radiography, and polar microscopy, the sorption component primarily contained iron?Cmagnesium (Fe?CMg) minerals in basalt and granite. Moreover, the distribution coefficient (K d) of Cs in various concentrations (~10?2?C10?7?M) obtained from batch tests indicated a higher sorption capacity in basalt than that in granite because of the 10% Fe?CMg mineral content. The diffusion of Cs in both crushed granite and basalt reached steady state after 110?days, and the apparent diffusion coefficients (D a) were 2.86?×?10?11 and 1.82?×?10?12?m2/s, respectively. However, the value of D a for Cs in intact rocks was estimated to be 1.45?×?10?12?m2/s in granite and 0.56?×?10?12?m2/s in basalt, lower than the values obtained using crushed rocks. In addition to the microporous structure (major sorption minerals), it showed that the major retardation of Cs depended on the porosity (??) of compacted media, according to through-diffusion tests. In fact, the solid/liquid (S/L) ratio decreased as is the case when switching from batch to column experiments and the sorption effect on minerals became even more negligible in retardation of radionuclide migration.  相似文献   

19.
Polycrystalline gold–nickel thin films are deposited on silicon (111) wafers by evaporation in a vacuum of 2 × 10?6 mbar. Concentration profiles of heat‐treated specimens are obtained by Auger electron depth profiling. The heat treatments are carried out in a vacuum furnace of 4 × 10?6 mbar in the temperature interval 473–773 K. The grain boundary diffusion coefficient is determined, using a modified Wipple model, to be (3 × 10?4 cm2 s?1) exp (?0.94 eV kT?1). It is concluded that interdiffusion in the investigated system is characterized by type B kinetics, and that grain boundary diffusion plays a dominant role in the mass transport process of such films. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

20.
The kinetics of the gas‐phase reactions of O3 with a series of selected terpenes has been investigated under flow‐tube conditions at a pressure of 100 mbar synthetic air at 295 ± 0.5 K. In the presence of a large excess of m‐xylene as an OH radical scavenger, rate coefficients k(O3+terpene) were obtained with a relative rate technique, (unit: cm3 molecule?1 s?1, errors represent 2σ): α‐pinene: (1.1 ± 0.2) × 10?16, 3Δ‐carene: (5.9 ± 1.0) × 10?17, limonene: (2.5 ± 0.3) × 10?16, myrcene: (4.8 ± 0.6) × 10?16, trans‐ocimene: (5.5 ± 0.8) × 10?16, terpinolene: (1.6 ± 0.4) × 10?15 and α‐terpinene: (1.5 ± 0.4) × 10?14. Absolute rate coefficients for the reaction of O3 with the used reference substances (2‐methyl‐2‐butene and 2,3‐dimethyl‐2‐butene) were measured in a stopped‐flow system at a pressure of 500 mbar synthetic air at 295 ± 2 K using FT‐IR spectroscopy, (unit: cm3 molecule?1 s?1, errors represent 2σ ): 2‐methyl‐2‐butene: (4.1 ± 0.5) × 10?16 and 2,3‐dimethyl‐2‐butene: (1.0 ± 0.2) × 10?15. In addition, OH radical yields were found to be 0.47 ± 0.04 for 2‐methyl‐2‐butene and 0.77 ± 0.04 for 2,3‐dimethyl‐2‐butene. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 394–403, 2002  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