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 23 毫秒
1.
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   

2.
A graft copolymer of polyisobutylene with poly(methyl methacrylate) [PMMA] was prepared containing naphthalene [N] groups statistically distributed along the PMMA backbone. Fluorescence studies of dilute solutions of the polymer in ethyl acetate showed small amounts of N-excimer emission when the NMA/MMA ratio was 1:9, and undetectable excimer when the ratio was 2/98. Energy transfer experiments were carried out between polymer-bound naphthalene and anthracene [A] added to the solution. The rate constant for energy transfer, kET, for the process N* + A → N + A* was found to be 1.1 × 1010 M−1 s−1 at 22°, slightly smaller than that (1.2 × 1010 M−1 s−1) for energy transfer from naphthylmethyl pivalate to A. From the value of the rate constants, a value for the segmental diffusion coefficient of polymer bound naphthalene was calculated.  相似文献   

3.
Diffusion of ions through hardened cement pastes   总被引:7,自引:0,他引:7  
Diffusion of Na+ and Cl ions through thin hardened paste section was determined as a function of water/cement ratio, curing time, curing temperature, and diffusion temperature. Diffusivities of Na+ ions were smaller than for Cl ions in the temperature range measured up to 60°C. Activation energies for diffusion were calculated, and contrasts were made with diffusion through a low porosity quartzite rock having minimal surface interaction with the diffusing species.  相似文献   

4.
We report that ultrastable faujasite-based ruthenium zeolites are highly active catalysts for N2O decomposition at low temperature (120–200°C). The faujasite-based ruthenium catalysts showed activity for the decomposition of N2O per Ru3+ cation equivalent to the ZSM-5 based ruthenium catalysts at much lower temperatures (TOF at 0.05 vol.-% N2O: 5.132 × 10−4 s−1 Ru−1 of Ru-HNaUSY at 200°C versus 5.609 × 10−4 s−1 Ru−1 of Ru-NaZSM-5 at 300°C). The kinetics of decomposition of N2O over a Ru-NaZSM-5 (Ru: 0.99 wt.-%), a Ru-HNaUSY (Ru: 1.45 wt.-%) and a Ru-free, Na-ZSM-5 catalyst were studied over the temperature range from 40 to 700°C using a temperature-programmed micro-reactor system. With partial pressures of N2O and O2 up to 0.5 vol.-% and 5 vol.-%, respectively, the decomposition rate data are represented by: −dN2O/dt=itk(PN2O) (PO2)−0.5 for Ru-HNaUSY, −dN2O/dt=k(PN2O) (PO2)−0.1 for Ru-NaZSM-5, and −dN2O/dt=k(PN2O)−0.2 (PO2)−0.1 for Na-ZSM-5. Oxygen had a stronger inhibition effect on the Ru-HNaUSY catalyst than on Ru-NaZSM-5. The oxygen inhibition effect was more pronounced at low temperature than at high temperature. We propose that the negative effect of oxygen on the rate of N2O decomposition over Ru-HNaUSY is stronger than Ru-NaZSM-5 because at the lower temperatures (<200°C) the desorption of oxygen is a rate-limiting step over the faujasite-based catalyst. The apparent activation energy for N2O decomposition in the absence of oxygen is much lower on Ru-HNaUSY (Ea: 46 kJ mol−1) than on Ru-NaZSM-5 (Ea: 220 kJ mol−1).  相似文献   

5.
This paper reports a comprehensive study by tast polarography, d.p.p., and cyclic voltammetry on the electrochemical reduction in different electrolytic media of ortho- and meta-nitrotoluene derivatives. Controlled potential electrolysis was used to generate the nitro radical anions and its detection by cyclic voltammetry and UV–Vis spectroscopy was performed. In protic media (30% ethanol/0.1 M Britton–Robinson buffer pH 2–12) both derivatives gave a sharp irreversible well-defined peak in all the pH range on Hg in a reaction involving four electrons to give the hydroxylamine derivative. In this medium meta-nitrotoluene is easier reduced in approximately 80 mV than that of the ortho-nitrotoluene. In mixed aqueous organic media (0.015 M aqueous citrate/DMF: 60:40, 0.3 M KCl and 0.1 M TBAI) at pH>8, the isolation and the electrochemical characterization of the one-electron reduction product, the nitro radical anion was achieved. At a 1 mM of nitrotoluene concentration, the average dismutation second-order rate constant values, k2, were: 11,000±170 and 6900±72 M−1 s−1 for ortho-and meta-nitrotoluene, respectively. In aprotic media (0.1 M TBAI in DMF), the nitro radical anions were more stable than that of mixed media, with the following dismutation second-order rate constant values, k2: 5800±35 and 4700±42 M−1 s−1 for ortho- and meta-nitrotoluene, respectively. Also, a comparison between nitrotoluene derivatives and some nitrocompounds of pharmacological relevance relating the effects of substituents on nitrobenzene and the electrolytic media composition on both the easiness of formation and stability of radicals is presented.  相似文献   

6.
8-Hydroxyquinoline-5-sulphonate/Al(III) aqueous solutions were studied both by potentiometric titrations and voltammetric measurements, in order to obtain the number, the stoichiometry and the stability constants of the complexes formed at equilibrium, and to evaluate the redox and (electro)kinetic properties of the free ligand and of the metal/ligand complexes. The complexes formed in 0.2 m (Na)Cl aqueous solution (stability log beta values ± standard deviation) are AlL+ (8.95 ± 0.05), AlL2 (17.43 ± 0.03) and AlL33− (24.58 ± 0.05), where “L” denotes the free ligand in the completely deprotonated form (L2−, pKa1 = 3.910 ± 0.008, pKa2 = 8.319 ± 0.004). AlL33− is the predominant Al(III) species in a very wide range of pH, metal and ligand concentrations and metal-to-ligand ratios. The free ligand shows an oxidation wave at 0.62 V versus SCE. The proposed oxidation mechanism includes a first reversible one-electron oxidation of the ligand, followed by a coupling reaction and by a second reversible one-electron oxidation, and finally by a decomposition reaction. The addition of Al(III) lowers the intensity of the oxidation wave due to the formation of the redox-inactive complex AlL33−. A residual low signal was attributed to the free ligand produced by the complex dissociation, AlL33− = AlL2 + L2−. All the kinetic parameters involved in the ligand oxidation and in the complex disruption were calculated on the basis of the agreement between experimental and simulated linear sweep and cyclic voltammetries. Correctness of the mechanisms proposed was further confirmed “a posteriori” by the agreement between potentiometric and linear sweep voltammetric results. The low residual signal observed in the presence of fully formed complex was attributed to the free ligand produced by the complex dissociation, having a kinetic constant estimated 0.2 s−1.  相似文献   

7.
Solar light photodegradation, catalysed by a pyrylium salt, of seven benzoic acids present in olive oil mill, has been studied. Significant percentages of photodegradation (20–40%) have been achieved after 6 h of solar exposure for six of the acids, even though they were expected to be difficult to oxidise, due to the presence of an electron-withdrawing carboxylic acid group directly attached to the aromatic ring. Quenching constants for the electron-transfer process between the substrate and the excited catalyst were calculated by means of fluorescence measurements for syringic acid (66×109 M−1 s−1), gallic acid (51×109 M−1 s−1), veratric acid (51×109 M−1 s−1), vanillic acid (48×109 M−1 s−1), protocatechuic acid (37×109 M−1 s−1) and p-hydroxybenzoic acid (15×109 M−1 s−1); no quenching was found for benzoic acid. These photophysical measurements are in good correlation with the yields obtained in the pyrylium salt photocatalysed degradation of those phenolic acids.  相似文献   

8.
The objective of this work was to study the promotional effect of Pt on Co-zeolite (viz. mordenite, ferrierite, ZSM-5 and Y-zeolite) and Co/Al2O3 on the selective catalytic reduction (SCR) of NOx with CH4 under dry and wet reaction stream. After being reduced in H2 at 350°C, the PtCo bimetallic zeolites showed higher NO to N2 conversion and selectivity than the monometallic samples, as well as a combination of the latter samples such as mechanical mixtures or two-stage catalysts. After the same pretreatment, under wet reaction stream, the bimetallic samples were also more active. Among the other catalysts studied with 5% of water in the feed, (NO = CH4 = 1000 ppm, O2 = 2%), the NO conversion dropped to zero over Co2.0Mor at 500°C and GHSV = 30,000 h−1, whereas it is 20% in Pt0.5Co2.0Mor. In Pt/Co/Al2O3 the NOx conversion dropped below 5% with only 2% of water under the same reaction conditions. The specific activity given as molecules of NO converted per total metal atom per second were 16.5 × 10−4 s−1 for Pt0.5Co2.0Fer, 13 × 10−4 s−1 for Pt0.5Co2.0Mor, 4.33 × 10−4 s−1 for Pt0.5Co2.0ZSM-5 and 0.5 × 10−4 s−1 for Pt/Co/Al2O3. The Y-zeolite-based samples were inactive in both mono and bimetallic samples. The species initially present in the solid were Pt° and Co°, together with Co2+ and Pt2+ at exchange positions. Co° seems not to participate as an active site in the SCR of NOx. Those species remained after the reaction but some reorganization occurred. A synergetic effect among the different species that enhances both the NO to NO2 reaction, the activation of CH4 and also the ability of the catalyst to adsorb NO, could be responsible for the high activity and selectivity of the bimetallic zeolites.  相似文献   

9.
Pdn+/Cen+/Na+/γ-Al2O3-type materials used as FCC additives for CO/NOx control were characterized by extended X-ray absorption fine structure (EXAFS) spectroscopy and in situ FTIR. The EXAFS data indicate that in freshly prepared samples palladium is present in the form of highly dispersed PdO species. Reduction with H2 at 500 °C leads to the formation of small Pd clusters incorporating on average approximately six to eight metal atoms at a Pd−Pd bond distance of 2.76 Å. All components of these materials can interact with NO and promote the formation of nitrate/nitrite species, essentially “trapping” NOx species on the catalyst surface. However, the Na+ species dominate the surface chemistry and readily form sodium nitrates with a characteristic IR band at 1370–1385 cm−1. Finally, hydroxyls from the support are also actively participating in the formation of HNOx type compounds with characteristic stretching vibrations in the 3500–3572 cm−1 region.  相似文献   

10.
The objective of this study was to evaluate the use of cheap carbon-based adsorbents from sewage sludge pyrolysis to remove Na+, K+, Ca2+ and Mg2+ from saline water. Four model solutions of NaCl, KCl, CaCl2 and MgCl2 that simulated seawater composition were used. The model solutions were 456.54 mmolc L−1 NaCl, 9.72 mmolc L−1 KCl, 19.96 mmolc L−1 CaCl2 and 111.09 mmolc L−1 MgCl2. Two carbon adsorbents, one with chemical activation, were prepared by a new method and evaluated for ion adsorption. The results indicated that carbon adsorbent without chemical activation was the most effective in removing ions from different solutions and the removal of metals followed the sequence: Na+ (76.78−69.66) >K+ (66.0−57.80) >Mg2+ (44.84−42.85) >Ca2+ (35.12−12.38). Results showed that it is to possible prepare carbon-based adsorbents from sewage sludge following inexpensive and environmentally acceptable methods.  相似文献   

11.
Measurements are presented of the compressive stress–strain behaviour of polycarbonate (PC) and polyvinylidene difluoride (PVDF) at strain rates from 10−4 to 104 s−1 at room temperature, and temperatures from −50 to +150 °C at 103 s−1. These results, obtained using a split Hopkinson pressure bar and Instron testing machine, are supported by dynamic mechanical analysis (DMA) measurements on the materials. Previous researchers have observed that the yield stress of these materials is bilinearly dependent on the logarithm of strain rate. The data presented here show that the bilinearity is due to the movement of low order transitions in the materials, so that they occur at temperatures above room temperature at the higher strain rates. In particular, these transitions are the β transition in PC, and the glass transition in PVDF. In addition, Appendix A presents measurements of a high strain rate Poisson's ratio of polycarbonate and its evolution with strain.  相似文献   

12.
-Alumina-supported MFI zeolite membranes were modified by on-stream catalytic thermal cracking of methyldiethoxysilane (MDES) molecules inside the zeolitic channels during the separation of H2/CO2 gas mixture at 450 °C and atmospheric pressure. The MDES vapor was carried by the H2/CO2 feed gas and the effect of modification was monitored continuously through online analysis of the permeate stream. The modified membrane exhibited a significant increase in H2 selectivity over CO2 with a moderate decrease in H2 permeance. At 450 °C, the modified MFI membrane obtained a H2/CO2 permselectivity of 17.5 with H2 single gas permeance of 1.86 × 10−7 mol m−2 s−1 Pa−1 as compared to a permselectivity of 2.78 and permeance of 2.75 × 10−7 mol m−2 s−1 Pa−1 for the membrane before modification. The modified membrane also showed good performance and stability in separation of H2/CO2 gas mixture containing up to 28.4% water vapor at 450 °C and atmospheric pressure.  相似文献   

13.
Rate data have been obtained for CO hydrogenation on a well-characterized 11.7% Co/TiO2 catalyst in a differential fixed bed reactor at 20 atm, 180–240°C, and 5% conversion over a range of reactant partial pressures. The resulting kinetic parameters can be used to model precisely and accurately the kinetics of this reaction within this range of conditions. Turnover frequencies and rate constants determined from this study are in very good to excellent agreement with those obtained in previous studies of other cobalt catalysts, when the data are normalized to the same conditions of temperature and partial pressures of the reactants. Based on this comparison CO conversion and the partial pressure of product water apparently have little effect on specific rate per catalytic site. The data of this study are fitted fairly well by a simple power law expression of the form −rCO=kPH20.74PCO−0.24, where k=5.1×10−3 s−1 at 200°C, P=10 atm, and H2/CO=2/1; however, they are best fitted by a simple Langmuir–Hinshelwood (LH) rate form −rCO=aPH20.74PCO/(1+bPCO)2 similar to that proposed by Yates and Satterfield.  相似文献   

14.
Medium energy (5–25 keV) 13C+ ion implantation into diamond (100) to a fluence ranging from 1016 cm−2 to 1018 cm−2 was performed for the study of diamond growth via the approach of ion beam implantation. The samples were characterized with Rutherford backscattering/channelling spectroscopy, Raman spectroscopy, X-ray photoemission spectroscopy and Auger electron spectroscopy. Extended defects are formed in the cascade collision volume during bombardment at high temperatures. Carbon incorporation indeed induces a volume growth but the diamond (100) samples receiving a fluence of 4 × 1017 to 2 × 1018 at. cm−2 (with a dose rate of 5 × 1015 at. cm−2 s−1 at 5 to 25 keV and 800 °C) showed no He-ion channelling. Common to these samples is that the top surface layer of a few nanometers has a substantial amount of graphite which can be removed by chemical etching. The rest of the grown layer is polycrystalline diamond with a very high density of extended defects.  相似文献   

15.
Electrochemical behavior of the water-soluble C60-γ-CD (1:2) inclusion complex has been studied on the hanging mercury drop electrode. A one-electron reversible adsorptive electro-reduction and three irreversible adsorptive electro-reductions were detected by cyclic voltammetry. The amount of C60-γ-CD adsorbed at saturation is 2.50 × 10−11 mol cm−2, the diffusion coefficient is 4.36 × 10−6cm2s−1 and the standard rate constant of the surface reaction ks are 0.745 s−1, 0.612 s−1 and 0.513s−1, respectively.  相似文献   

16.
17.
The mineral matter in an Australian black coal has been isolated using a low-temperature ashing (LTA) procedure. This LTA procedure is a modification of the Australian Standard for LTA at 370°C, and alleviates adverse effects to the minerals caused by the heat of combustion. The leaching behaviour of the mineral matter towards aqueous HCl and hydrofluoric acid (HF) is presented. HCl can dissolve simple compounds such as phosphates and carbonates, yet it cannot completely dissolve the clays. HF reacts with almost every mineral in the mineral matter, except pyrite, and most of the reaction products are water soluble. However, at HF concentrations greater than that required to dissolve the aluminosilicate compounds in the mineral matter, insoluble compounds form. These compounds include CaF2, MgF2 and a compound containing Na, which is believed to be NaAlF4. It is proposed that HF reacts preferentially with the aluminosilicates in the mineral matter to form largely AlF2+, AlF3 and SiF4, and that the concentrations of free fluoride (F) and AlF4 are not high enough to complex cations such as Ca2+, Mg2+ and Na+. When the mineral matter is treated with HF concentrations greater than that required to dissolve all of the aluminosilicates, AlF3, AlF4 and SiF62− form, the concentration of F is high enough to complex Ca2+ and Mg2+ and form insoluble CaF2 and MgF2, and the concentration of AlF4 is high enough to complex Na+ and form insoluble NaAlF4. This work has application toward the development of a process for producing Ultra Clean Coal with less than 0.1% by weight mineral matter.  相似文献   

18.
It is well-known that the electrolytes can influence the electrochemical reduction of carbon dioxide (ERCO2) in aqueous media. In this work, we explore the effects of alkali metal cations and anions (Li+, Na+, K+, Rb+, Cs+, HCO3-, Cl-, Br-, I-) on the current density and product selectivity for the ERCO2 into formic acid (HCOOH) on the SnO2/carbon paper (SnO2/C) electrode. Results of the ERCO2 experiments show that for the cations, the promotion effects on current density and faradaic efficiencies (FEs) are in the order of Li+ < Na+ < K+ < Cs+ < Rb+. For the anions, the current density values are in the order of NaHCO3 < NaCl < NaBr < NaI and KHCO3 < KCl ≈ KI < KBr, respectively, and that on the FEs for the formation of the HCOOH (FEHCOOH) is HCO3- < Cl- < Br- < I-. Based on this result, the effects of alkali metal cations and anions on ERCO2 are discussed.  相似文献   

19.
The electrochemistry of cesium was investigated at mercury electrodes in the tri-1-butylmethylammonium bis((trifluoromethyl)sulfonyl)imide (Bu3MeN+Tf2N) room-temperature ionic liquid (RTIL) by using cyclic staircase voltammetry, rotating disk electrode voltammetry, and chronoamperometry. The reduction of cesium ions at mercury exhibits quasireversible behavior with k0 = 9.8 × 10−5 cm s−1 and = 0.36. The diffusion coefficient of Cs+ in this RTIL was 1.04 × 10−8 cm2 s−1 at 303 K. Bulk deposition/stripping experiments conducted at a rotating mercury film electrode gave an average recovery of 97% of the electrodeposited Cs. The density, absolute viscosity, and equivalent conductance of Bu3MeN+Tf2N were measured over the range of temperatures from 298 to 353 K. A polynomial equation describing the temperature dependence of the density is presented. Both the viscosity and conductance exhibited the non-Arrhenius temperature dependence typical of glass-forming liquids. The ideal glass transition temperature and the activation energies for viscosity and conductance were obtained by fitting the Vogel–Tammann–Fulcher (VTF) equation to the experimental data for these transport properties.  相似文献   

20.
A novel facilitated transport membrane for gas separation using a capillary membrane module is proposed in which a carrier solution is forced to permeate the membrane. Both a feed gas and a carrier solution are supplied to the lumen side (high pressure side, feed side) of the capillary ultrafiltration membrane and flow upward. Most of the carrier solution which contains dissolved solute gas, CO2 in the present case, permeates the membrane to the permeate side (low pressure side, shell side), where the solution liberates dissolved gas to form a lean solution. The lean solution is circulated to the lumen side. This type of capillary membrane module was applied to the separation of CO2 from model flue gases consisting of CO2 and N2. Monoethanolamine (MEA), diethanolamine (DEA) and 2-amino-2-methyl-1-propanol (AMP) were used as carriers or absorbents of CO2. The feed side pressure was atmospheric and the permeate side was evacuated at about 10 kPa. CO2 in the feed gas was successfully concentrated from 5–15% to more than 98%. The CO2 permeance was as high as 2.7×10−4 mol m−2 s−1 kPa−1 (8.0×10−4 cm3 cm−2 s−1 cmHg−1) when the CO2 mole fraction in the feed was 0.1 and temperature was 333 K. The selectivity of CO2 over N2 was in the range from 430 to 1790. The membrane was very stable over a discontinuous one-month testing period.  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