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 standard molal potentials E°m of the Hg/Hg2(OPr)2, OPr electrode at 15°, 20°, 25°, 30° and 35° C have been determined. The E°m values obtained are 0.5114, 0.5072, 0.5031, 0.4988 and 0.4942 V respectively, which can be fitted to the equation Edgm/V = 0.5031 −8.56 × 10−4 (itt/°C − 25)−3.0588 × 10−6 (t/ °C -25)2. The changes in standard free energy, entropy and enthalpy for the cell reaction have been calculated.  相似文献   

2.
Layered -titanate materials, NaxMx/2Ti1−x/2O2 (M=Co, Ni and Fe, x=0.2–0.4), were synthesized by flux reactions, and electrical properties of polycrystalline products were measured at 300–800 °C. After sintering at 1250 °C in Ar, all products show n-type thermoelectric behavior. The values of both d.c. conductivity and Seebeck coefficient of polycrystalline Na0.4Ni0.2Ti0.8O2 were ca. 7×103 S/m and ca. −193 μV/K around 700 °C, respectively. The measured thermal conductivity of layered -titanate materials has lower value than conductive oxide materials. It was ca. 1.5 Wm−1 K−1 at 800 °C. The estimated thermoelectric figure-of-merit, Z, of Na0.4Ni0.2Ti0.8O2 and Na0.4Co0.2Ti0.8O2 was about 1.9×10−4 and 1.2×10−4 K−1 around 700 °C, respectively.  相似文献   

3.
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   

4.
Maria Andrei  Massimo Soprani 《Polymer》1998,39(26):7041-7047
A new class of polymer electrolytes, based on the interpenetrating polymer network approach, was obtained starting from functionalised macromers, of poly-ether nature, in the presence of a lithium salt (LiBF4, LiClO4, LiCF3SO3) and propylene carbonate (PC) or tetraethyleneglycol dimethylether (TGME), as plasticizers.

The macromers were synthesised by living polymerisation employing a HI/I2 system as the initiator. The macromer has a polymerisable end group, which can undergo radical polymerisation, attached to a monodisperse poly-vinylether, containing suitable ethylene oxide groups for ion coordination. Monomers and macromers were characterised by FTi.r., u.v.–vis, 1H- and 13C-n.m.r.

Self-consistent and easily handled membranes were obtained as thin films by a dry procedure using u.v. radiation to polymerise and crosslink the network precursors, directly on suitable substrates, in the presence of the plasticizer and the lithium salt. The electrolytic membranes were studied by complex impedance and their thermal properties determined by differential scanning calorimetry analysis.

Ionic conductivities (σ) were measured for PC and TGME-based membranes at various plasticizer and salt contents as a function of T (60 to −20°C). LiClO4/PC/PE electrolytes, with 3.8% (w/w) salt and 63% PC, have the highest σ (1.15×10−3 and 3.54×10−4 S cm−1 at 20°C and −20°C, respectively). One order of magnitude lower conductivities are achieved with TGME; samples with 6% (w/w) LiClO4 and 45% (w/w) TGME exhibit σ values of 2.7×10−4 and 2.45×10−5 S cm−1 at 20°C and −20°C.  相似文献   


5.
The crystal structures of fully dehydrated Sr46–X [Sr46Si100Al92O384; a=25.214(7) Å] and of its ammonia sorption complex, Sr46–X·102NH3 [Sr46Si100Al92O384·102NH3; a=25.127(7) Å], have been determined by single-crystal X-ray diffraction techniques in the cubic space group Fd at 21(1)°C. The Sr46–X crystal was prepared by ion exchange in a flowing stream of aqueous 0.05 M Sr(ClO4)2 for 5 days followed by dehydration at 360°C and 2×10−6 Torr for 2 days. To prepare the ammonia sorption complex, another dehydrated Sr46–X crystal was exposed to 230 Torr of zeolitically dried ammonia gas for 1 h followed by evacuation for 12 h at 21(1)°C and 5×10−4 Torr. The structures were refined to the final error indices, R1=0.043 and Rw=0.039 with 466 reflections, and R1=0.049 and Rw=0.044 with 382 reflections, for which I>3σ(I). In dehydrated Sr46–X, all Sr2+ ions are located at two crystallographic sites. 16 Sr2+ ions are at the centers of the double six-rings, filling that site (site I, Sr–O=2.592(6) Å). The remaining 30 Sr2+ ions are in the supercage (site II); each extends 0.56 Å into the supercage from the plane of its three nearest oxygen atoms (Sr–O=2.469(6) Å). In the structure of Sr46–X·102NH3, the Sr2+ ions are located at three crystallographic sites: 12 are found at site I [Sr–O=2.652(10) Å]; four in the sodalite units (site I′) each coordinated to three framework oxygen atoms at 2.654(9) Å and also to three ammonia molecules at 2.76(8) Å. The remaining 30 Sr2+ ions lie at site II. Each extends 1.12 Å into the supercage where it coordinates to three framework oxygen atoms at 2.584(7) Å and also to three ammonia molecules at 2.774(24) Å.  相似文献   

6.
The polarographic behaviour of methyl-3-mercaptopropionate, H3COOC CH2CH2SH has been studied in 25% ethanol, 0·1 M NaClO4 and 0·002% Triton X-100 in the pH range 4·60−11·36 at a dme. The electrode process is irreversible and the plateau current is diffusion-controlled. The half wave potentials of the waves shift towards more negative potentials with increase in pH and the dissociation constant of the sulphydryl group is found to be 9·95. The values of the kinetic parameters, transfer coefficient () and formal rate constant (K°b,h), have been calculated by Koutecky's method at pH 9·37 and are 0·48 and 1·13 × 10−4 cm/s respectively.  相似文献   

7.
We synthesized high-quality and oriented periodic mesoporous organosilica (PMO) monoliths through a solvent evaporation process using a wide range of mole ratios of the components: 0.17–0.56 1,2-bis(triethoxysilyl)ethane (BTSE): 0.2 cetyltrimethylammonium chloride (CTACl): 0–1.8 × 10−3 HCl: 0–80 EtOH: 5–400 H2O. X-ray diffraction (XRD) patterns and transmission electron microscopy (TEM) images indicated that the mesoporous channels within the monolith samples were oriented parallel to the flat external surface of the PMO monolith and possessed a hexagonal symmetry lattice (p6mm). The PMO monolith synthesized from a reactant composition of 0.35 BTSE: 0.2 CTACl: 1.8 × 10−6 HCl: 10 EtOH: 10 H2O had a pore diameter, pore volume, and surface area – obtained from an N2 sorption isotherm – of 25.0 Å, 0.96 cm3 g−1 and 1231 m2 g−1, respectively. After calcination at 280 °C for 2 h in N2 flow, the PMO monolith retained monolith-shape and mesostructure. Pore diameter and surface area of the calcined PMO monolith sample were 19.8 Å, 0.53 cm3 g−1 and 1368 m2 g−1, respectively. We performed 29Si and 13C CP MAS NMR spectroscopy experiments to confirm the presence of Si–C bonding within the framework of the PMO monoliths. We investigated the thermal stability of the PMO monoliths through thermogravimetric analysis (TGA). In addition, rare-earth ions (Eu3+, Tb3+ and Tm3+) were doped into the monoliths. Optical properties of those Eu3+, Tb3+ and Tm3+-doped PMO monoliths were investigated by photoluminescence (PL) spectra to evaluate their potential applicability as UV sensors.  相似文献   

8.
The hydrodynamic characteristics in aqueous solution at ionic strength I=0.2  of carboxymethylchitins of different degrees of chemical substitution have been determined. Experimental values varied over the following ranges: the translational diffusion coefficient (at 25.0°C), 1.1<107×D<2.9 cm2 s−1; the sedimentation coefficient, 2.4<s<5.0 S; the Gralen coefficient (sedimentation concentration-dependence parameter), 130<ks<680 mL g−1; the intrinsic viscosity, 130<[η]<550 mL g−1. Combination of s with D using the Svedberg equation yielded ‘sedimentation–diffusion' molecular weights in the range 40 000<M<240 000 g mol−1. The corresponding Mark–Houwink–Kuhn–Sakurada (MHKS) relationships between the molecular weight and s, D and [η] were: [η]=5.58×10−3 M0.94; D=1.87×10−4 M−0.60; s=4.10×10−15 M0.39. The equilibrium rigidity and hydrodynamic diameter of the carboxymethylchitin polymer chain is also investigated on the basis of wormlike coil theory without excluded volume effects. The significance of the Gralen ks values for these substances is discussed.  相似文献   

9.
The structure of a cyclopropane sorption complex of dehydrated fully Mn2+-exchanged zeolite X, Mn46Si100Al92O384 · 30C3H6 (a=24.690(4) Å), has been determined by single-crystal X-ray diffraction techniques in the cubic space group at 21(1)°C. The crystal was prepared by ion exchange in a flowing stream of 0.05 M aqueous Mn(NO3)2 for three days, followed by dehydration at 460°C and 2×10−6 Torr for two days, and exposure to 100 Torr of cyclopropane gas at 21(1)°C. The structure was determined in this atmosphere and was refined to the final error indices R1=0.065 and R2=0.071 with 509 reflections for which I>3σ (I). In this structure, Mn2+ ions are located at two crystallographic sites. Sixteen Mn2+ ions fill the octahedral site I at the centers of the hexagonal prisms (Mn–O=2.290(9) Å). The remaining 30 Mn2+ ions are at site II; each extends 0.41 Å into the supercage (an increase of 0.27 Å upon C3H6 sorption as compared to fully dehydrated Mn46Si100Al92O384) where it coordinates to three trigonally arranged framework oxygens at 2.148(8) Å and complexes weakly and facially to a cyclopropane molecule by a primarily quadrupolar interaction. The carbon atoms of each cyclopropane molecule are equivalent and equidistant from its Mn2+ ion (Mn–C=2.95(9) Å). Because of high thermal motion, the C–C bond length is inaccurately determined; the value found, 1.21(8) Å, is too small.  相似文献   

10.
Preparations and physico-chemical characterizations of NASICON-type compounds in the system Li1+xAlxA2−xIV(PO4)3 (AIV=Ti or Ge) are described. Ceramics have been fabricated by sol-gel and co-grinding processes for use as ionosensitive membrane for Li+ selective electrodes. The structural and electrical characteristics of the pellets have been examined. Solid solutions are obtained with Al/Ti and Al/Ge substitutions in the range 0≤x≤0·6. A minimum of the rhombohedral c parameter appears for x about 0·1 for both solutions. The grain ionic conductivity has been characterized only in the case of Ge-based compounds. It is related to the carrier concentration and the structural properties of the NASICON covalent skeleton. The results confirm that the Ti-based framework is more calibrated to Li+ migration than the Ge-based one. A grain conductivity of 10−3 S cm−1 is obtained at 25°C in the case of Li1·3Al0·3Ti1·7(PO4)3. A total conductivity of about 6×10−5 S cm−1 is measured on sintered pellets because of grain boundary effects. The use of such ceramics in ISE devices has shown that the most confined unit cell (i.e. in Ge-based materials) is more appropriate for selectivity effect, although it is less conductive.©  相似文献   

11.
J. Hank  H. Hoff 《Electrochimica acta》1972,17(12):2363-2369
In measuring the Hall effect in aqueous solutions some aspects of electrochemistry and electrode kinetics have to be considered. The described apparatus is based on a constant magnetic field and an alternating electrical field of variable frequency and on three Hall electrodes. The hall mobility in 0·1 N HNO3, 0·01 N AgNO3 at 20°C is 2·58 × 10−3 cm/Vs.

Résumé

Quelques aspects électrochimiques et électrocinétiques des mesures de l'effet Hall dans les solutions aqueuses sont discutés. L'apparaillage décrit en détail fonctionne au moyen d'un champ magnétique continu, d'un champ électrique alternatif de fréquence variable et avec trois sondes d'Hall. La mobilité d'Hall trouvée dans un électrolyte de 1/10 n HNO3, 1/100 n AgNO3 est = 2,58 · 10−3 cm2/Vs à 20°C.  相似文献   


12.
A series of ceramics samples, Na5+xYAlxSi4-xO12, has been prepared by a solid state reaction with the starting materials of SiO2, Y2O3, Al2O3 and Na2CO3. Their crystalline structure and morphology have been studied by the determination of XRD, IR, TG, DTA and SEM. Their conductivity has been measured by means of the complex impedance method. The dependence of the conductivity and density of the samples on the amount of the added Al2O3 and the reaction between the conductivity and the temperature have been discussed. When x = 0, the density of the sintering sample is 90% T.D., and the conductivity is 1·48 x 10−1 (ωcm)−1 at 300°C; when x = 0·1, the density is up to 97% T.D., and the conductivity up to 1·74 x 10−1 (ω cm)−1 at 300°C.  相似文献   

13.
Phase transition phenomenon of the 1-dodecanol monolayer at the air/water interface was studied by the dynamic γ(t) curves and the adsorption isotherm obtained by ellipsometry at 20 °C. The surface-concentration adsorption isotherm clearly showed three abrupt increases at bulk concentration C of 1.3 × 10−9, 2 × 10−9 and 3.7 × 10−9 mol/mL, respectively. The 1st and the 3rd transitions observed herein, that were typical 2D first-order transitions, were consistent with the gas to liquid expanded (G–LE) and the liquid expanded to liquid condensed (LE–LC) phase transitions observed in a previous tensiometry study. The 2nd transition that occurred at C = 2 × 10−9 mol/mL was not identified from any previous dynamic surface-tension profiles. Judging from the substantial increase in the film thickness of the transition, it was believed that the orientation change of the adsorbed molecule was involved in the LE phase. A LEh and a LEv phase, that denoted the “lie-down” and “stand-up” types of adsorption, respectively, was used to describe this transition and a cusp, instead of a constant surface-tension region, was observed in the dynamic γ(t) curves for this transition. This suggested that, since the surface tension varied during the transition process, the newly identified LEh and LEv transition might not be the typical first-order type of phase transition.  相似文献   

14.
The electrochemical reduction of oxygen in 0.88 M KOH at 25°C was studied using two silver plated copper electrodes that contained uniform, parallel, tapered pores. Photolithography, etching, and plating techniques were used to fabricate these electrodes. Exchange current densities of 2 × 10−5 A cm−2 and 2 × 10−6 A cm−2 were measured using an initial slope method. Diffusion limiting currents increased with differential pressure across the gas-electrolyte interface which suggests smaller contact angles at higher pressures. A wedge meniscus model that treats activation, ohmic, and concentration overpotential was applied to fit the polarization curves.  相似文献   

15.
A new brownmillerite-related compound. Ca2Cr2O5, has been prepared. It has been indexed according to an orthorhombic lattice a = 5·750 Å, b = 14·398 Å and c = 5·483 Å. A series of experiments was performed in order to find the appropriate firing temperature. The total conductivity was measured by a four-point method in the range of 690–911°C. Impedance spectroscopy was also employed in the temperature range 343–785°C. Conductivity measurements at different oxygen pressures at 500°C suggest that Ca2Cr2O5 is a predominantly ionic conductor at Po2 = 1–10−2atm.  相似文献   

16.
The electrochemical reduction of the zinc ion to the amalgam has been studied in lithium perchlorate-dimethylformamide solution. The true standard rate constant in 2·5 × 10−3 cm sec and the cathodic charge transfer coefficient is 0·66.  相似文献   

17.
Boshan Yu  Quan Li 《Desalination》1993,90(1-3):335-344
This paper reports on a new technology for removing chlorine from dilute aqueous solutions produced in electrolysis of potassium chloride using a flat Teflon microporous membrane. Preliminary experiments indicate that at 55°C the overall mass transfer coefficient is 5.5 × 10−5 m/s. The treatment capacity of 2.78 × 10−5 m3/m2· s was examined. This technology is superior to conventional air blow-out methods and has considerable commercial potential.  相似文献   

18.
The calculations of chemical equilibria for the Zr---B---N system using the VCS algorithm were carried out for a temperature range of 27–2127°C and 1·3 × 10−3 −1 × 107 Pa pressure range. Two BN : Zr molar ratios were selected: 1 : 1 and 2 : 1. It follows from these studies thatZr can react with BN in a very wide range of pressures and temperatures giving two new phases, namely ZrB2 and ZrN. The formation of these phases is strongly influenced by pressure-temperature parameters. In order to verify theoretical calculations by experiment, equimolar and 2 : 1 BN : Zr batches were prepared and heated at 1427°C and 1627°C in vacuum and in argon. Phase identification by X-ray diffraction fully corroborated the computational results.  相似文献   

19.
Impedance measurements have been performed in molten PbCl2—KCl at 400°C to investigate the mechanism of In(III) reduction and sulphide ion oxidation. In(III) reduction is purely diffusion controlled and represented by a Warburg impedance. The diffusion coefficient obtained for In(III) is D = 0.49 × 10−5 cm2s−1. The first oxidation step of sulphide is rapid and also diffusion controlled.  相似文献   

20.
Some novel PVC membrane electrodes based on N,N′-bis-(benzaldehyde)-glycine metallic complexes of Cu(II), Ni(II), Zn(II) and Co(II) as neutral carriers are described. The results showed that the electrode based on the N,N′-bis-(benzaldehyde)-glycine copper(II) complex [Cu(II)-BBAG] had a near-Nernstian response to the thiocyanate ion ranging from 1.0×10−1 to 9.0×10−7 M in a phosphate buffer solution of pH 4.0 with a detection limit of 7.0×10−7 M and a slope of −57.6 mV/decade at 25°C. The electrode displays an anti-Hofmeister selectivity sequence in the following order: SCN>ClO4>Sal>I>Br>NO2>NO3>SO2−3>SO2−4>Cl>H2PO The response mechanism is discussed in view of both AC impedance and UV spectroscopy techniques. The [Cu(II)-BBAG]-based electrode was successfully applied to the determination of the thiocyanate ion in wastewater and human saliva.  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