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.
Fe-10Cr nanocrystalline (nc) coatings with a grain size of 20-30 nm were synthesized on glass substrates by magnetron sputtering. The corrosion behavior was investigated in 0.05 mol/L H2SO4 + 0.25 mol/L Na2SO4 and 0.05 mol/L H2SO4 + 0.5 mol/L NaCl solution by polarization curves, EIS and Mott-Schottky analysis. The results showed that compared to Fe-10Cr cast alloy, the active dissolution of the coating was accelerated; the passive film contained more Cr and therefore the coating was easier to passivate. The passive films formed on Fe-10Cr nc and cast alloy exhibited n-type semiconducting behavior in acidic solutions without Cl and p-type semiconducting behavior in acidic solutions with Cl. The lower breakdown potential for both materials in the solution with Cl is related to the p-type passive film formed on them. For Fe-10Cr nc, lower donor density and increased Cr content were responsible for the chemical stability of the passive film.  相似文献   

2.
The dissolution kinetics in 2 M H2SO4 of variously dehydroxylated nickeliferous goethites was investigated for five oxide-type lateritic nickel deposits. Goethite was the main constituent with minor amounts of quartz, talc, kaolinite and Mn oxides. Dissolution of Fe from heated materials followed the Kabai equation. There was a 9–34-fold increase in the Kabai dissolution rate constant (k) for samples heated at 340–400 °C due to both the increased surface area (1.5–2.6 fold) and higher density of structural defects (5–10 fold) in the variously dehydroxylated products. The presence of structural Al and Cr in goethite appears to reduce dissolution rate possibly through the greater M3+–OH, O bond strength relative to Fe3+, Ni2+–OH, O. Nickel showed congruent dissolution with Fe indicating that Ni was uniformly incorporated in the goethite structure. Pre-heating goethite to 600–800 °C for 30 min resulted in incongruent dissolution of Fe and Ni. It is postulated that some Ni is ejected from the neo-formed hematite structure and resides on the crystal surface or in voids. These results may contribute to the development of more efficient procedures for Ni extraction including heap leaching of lateritic nickel ores.  相似文献   

3.
Grain legumes are important crops, but they are salt sensitive. This research dissected the responses of four (sub)tropical grain legumes to ionic components (Na+ and/or Cl) of salt stress. Soybean, mungbean, cowpea, and common bean were subjected to NaCl, Na+ salts (without Cl), Cl salts (without Na+), and a “high cation” negative control for 57 days. Growth, leaf gas exchange, and tissue ion concentrations were assessed at different growing stages. For soybean, NaCl and Na+ salts impaired seed dry mass (30% of control), more so than Cl salts (60% of control). All treatments impaired mungbean growth, with NaCl and Cl salt treatments affecting seed dry mass the most (2% of control). For cowpea, NaCl had the greatest adverse impact on seed dry mass (20% of control), while Na+ salts and Cl salts had similar intermediate effects (~45% of control). For common bean, NaCl had the greatest adverse effect on seed dry mass (4% of control), while Na+ salts and Cl salts impaired seed dry mass to a lesser extent (~45% of control). NaCl and Na+ salts (without Cl) affected the photosynthesis (Pn) of soybean more than Cl salts (without Na+) (50% of control), while the reverse was true for mungbean. Na+ salts (without Cl), Cl salts (without Na+), and NaCl had similar adverse effects on Pn of cowpea and common bean (~70% of control). In conclusion, salt sensitivity is predominantly determined by Na+ toxicity in soybean, Cl toxicity in mungbean, and both Na+ and Cl toxicity in cowpea and common bean.  相似文献   

4.
The efficiency of the ‘ferrite process’ for the purification of wastewater heavily contaminated with nickel is evaluated, and the solid residues formed are characterised. The efficiency of the purification process is always above 99.9% for Fe2+/Ni2+ ratios greater than 3. The tested Fe2+/Ni2+ molar ratios (15/1, 7/1 and 3/1) yielded three different nickel ferrites. Inductively-coupled plasma atomic emission spectroscopy (ICP-AES), potentiometric titration, X-ray fluorescence (XRF), X-ray diffractometry (XRD) and differential scanning calorimetry (DSC) yielded NixFe1−xIIFe2IIIO4 (x=0.18, 0.40 and 0.65, respectively) as the most probable stoichiometry, and inverse spinel as the most probable structure. Heating at 600 °C causes the transformation of the solids into a mixture of NiFe2O4, α-Fe2O3 and NiO. Electrochemical analysis of the solid nickel ferrites was performed using carbon paste electrodes (CPEs) in HClO4 and HCl media. In each case, the first cyclic voltammogram showed the participation of solid species in the electrochemical transformation process, since the shape of the redox peaks could be related to the structure and stoichiometry of the ferrites. In second and successive scans, the voltammograms indicated the redox couples Feads3++1e⇔Feads2+ (0.525 V vs. Ag/AgCl) and Niads2++2e⇔Ni(s) (−0.470 V) in HClO4, and FeCl2,ads++1e⇔FeClads++Cl (0.475 V) and NiClx,ads(x−2)−+2e⇔Ni(s)+xCl (−0.550 V) in HCl.  相似文献   

5.
The electrochemistry of molten LiOH–NaOH, LiOH–KOH, and NaOH–KOH was investigated using platinum, palladium, nickel, silver, aluminum and other electrodes. The fast kinetics of the Ag+/Ag electrode reaction suggests its use as a reference electrode in molten hydroxides. The key equilibrium reaction in each of these melts is 2 OH = H2O + O2– where H2O is the Lux-Flood acid (oxide ion acceptor) and O2– is the Lux–Flood base. This reaction dictates the minimum H2O content attainable in the melt. Extensive heating at 500 °C simply converts more of the alkali metal hydroxide into the corresponding oxide, that is, Li2O, Na2O or K2O. Thermodynamic calculations suggest that Li2O acts as a Lux–Flood acid in molten NaOH–KOH via the dissolution reaction Li2O(s) + 2 OH = 2 LiO + H2O whereas Na2O acts as a Lux–Flood base, Na2O(s) = 2 Na+ + O2–. The dominant limiting anodic reaction on platinum in all three melts is the oxidation of OH to yield oxygen, that is 2 OH 1/2 O2 + H2O + 2 e. The limiting cathodic reaction in these melts is the reduction of water in acidic melts ([H2O] [O2–]) and the reduction of Na+ or K+ in basic melts. The direct reduction of OH to hydrogen and O2– is thermodynamically impossible in molten hydroxides. The electrostability window for thermal battery applications in molten hydroxides at 250–300 °C is 1.5 V in acidic melts and 2.5 V in basic melts. The use of aluminum substrates could possibly extend this window to 3 V or higher. Preliminary tests of the Li–Fe (LAN) anode in molten LiOH–KOH and NaOH–KOH show that this anode is not stable in these melts at acidic conditions. The presence of superoxide ions in these acidic melts likely contributes to this instability of lithium anodes. Thermal battery development using molten hydroxides will likely require less active anode materials such as Li–Al alloys or the use of more basic melts. It is well established that sodium metal is both soluble and stable in basic NaOH–KOH melts and has been used as a reference electrode for this system.  相似文献   

6.
Cyclic voltammetric experiments were carried out on platinum in acidic solution (pH 3) containing ferrous sulfate, nickel sulfate and ethylamines (EtNH2, Et2NH, Et3N). Spectral ultraviolet absorption studies indicate the complexation of both Fe2+ and Ni2+ ions with ethylamines. The results under transient polarisation conditions indicate the reduction of Fe2+ ions through the intermediate species FeOH+, with second electron transfer as a slow step. The higher charge transfer rate of FeOH+ over NiOH+ reduction causes the anomalous codeposition of Fe–Ni alloy film. Among the ethylamines, Et3N considerably assists the alloy deposition process. A gradual variation in free energy of alloy formation with Fe2+:Ni2+ (mol:mol) in the bath suggests the formation of an alloy intermediate phase rich in iron. Stripping voltammetric curves indicate the preferential dissolution of iron from iron rich alloy intermediate phase. X-ray diffraction studies further confirm the phase to be b.c.c. Fe–Ni alloy. The extent of corrosion of the Fe–Ni alloy film in the presence of ethylamines is in the following order: Et3N > Et2NH > EtNH2.  相似文献   

7.
《Ceramics International》2017,43(13):9960-9967
P2-type layered Na2/3Ni1/4Mn3/4O2 has been synthesized by a solid-state method and its electrochemical behavior has been investigated as a potential cathode material in aqueous hybrid sodium/lithium ion electrolyte by adopting activated carbon as the counter electrode. The results indicate that the Na+/Li+ ratio in aqueous electrolyte has a strong influence on the capacity and cyclic stability of the Na2/3Ni1/4Mn3/4O2 electrode. Increase on the Li+ content leads to a shift of the redox potential towards a high value, which is favorable for the improvement of the working voltage of the layered material as cathode. It is found that the coexistence of Na+ and Li+ in aqueous electrolyte can improve the cyclic stability for the Na2/3Ni1/4Mn3/4O2 electrode. A reversible capacity of 54 mAh g−1 was obtained with a high cyclability as the Na+/Li+ ratio was 2:2. Furthermore, an aqueous hybrid ion cell was assembled with the as-proposed Na2/3Ni1/4Mn3/4O2 as cathode and NaTi2(PO4)3/graphite synthesized in this work as anode in 1 M Na2SO4/Li2SO4 (mole ratio as 2:2) mixed electrolyte. The cell shows an average discharge voltage at 1.2 V, delivering an energy density of 36 Wh kg−1 at a power density of 16 W kg−1 based on the total mass of the active materials.  相似文献   

8.
Effect of sulphate anions on tunnel etching of aluminium   总被引:2,自引:0,他引:2  
Electrochemical etching of hard aluminium foil was studied at 100°C in NaCl solutions without and with Na2SO4 in concentrations up to 1.0m. Addition of Na2SO4 resulted in an increase in electric capacitance, in refinement of etch configuration and in an increase in tunnel density per unit volume. A decrease in the number of pits from which the tunnels grew also occurred. The capacitance increased with increasing concentration of Na2SO4 up to about 0.351 m, and then decreased. Sulphate ions depressed the formation of pits on the outer oxide-covered surface, but enhanced the growth of the pits and the formation of tunnels from the pits. It is suggested that the retardation of pit nucleation and the acceleration of tunnel growth in the presence of SO 4 2– ions can be explained by a partial replacement of Cl ions from the oxide and metal surface, respectively. Smaller diameter tunnels may be due to the formation of Al2(SO4)3 which can, in part, replace more aggressive AlCl3, and to an easier formation of a passivating film on the tunnel walls owing to their slower dissolution in the presence of Al2(SO4)3.  相似文献   

9.
The temperature–concentration dependences of the electrical conductivity and the activation energy for electrical conduction of glasses in the Na2O–B2O3 and Na2O–2PbO · B2O3 systems are studied. The investigation into the nature of the electrical conduction in these glasses reveals that the contribution from the electronic component (10–3%) of the conductivity is within the sensitivity of the Liang–Wagner technique. A considerable alkali conductivity is observed upon introduction of more than 12 mol % Na2O. The true transport number of sodium Na is as large as unity at [Na2O] 15 mol %. It is shown that the observed temperature–concentration dependences of the electrical and transport properties are governed by the ratio between the concentrations of polar and nonpolar structural–chemical units of the Na+[BO4/2], Na+[OBO2/2] Na+[OBO2/2], Pb2+ 1/2[BO4/2], Pb2+ 1/2[OBO2/2], and [BO3/2] types.  相似文献   

10.
The anodic behaviour of four cast iron alloys containing up to 16.7% Ni, in deaerated 60 wt% H3PO4 with and without 5 × 10–3 M F, Cl ions and 1:1 Cl/F mixture was studied by the potentiostatic technique. Values of E corr of the alloys are markedly influenced by their composition. The anodic behaviour in the active region is controlled by Fe in the alloys and the dissolution reaction is characterized by Tafel slopes, b a, between 64 and 88 mV (decade)–1. A two-electron transfer mechanism for the anodic dissolution is proposed. Passivation of the alloys is due to the formation of oxide layers including Fe2O3 and/or Fe3O4. Both critical and passive c.d. (I cc and I P) are markedly increased in the presence of Cl ions, but the presence of F ions inhibit the active dissolution of the alloys. The Tafel slope for oxygen evolution reaction (o.e.r.) in the transpassive region, is 240 ± 25 mV. In the proposed mechanism for the o.e.r., the rate determining step is an electron transfer reaction and possible interpretation of the high Tafel slopes is given based on the dual barrier model.  相似文献   

11.
Dissolution mechanism of colemanite in sulphuric acid solutions   总被引:1,自引:0,他引:1  
Boron compounds are very important raw materials in many branches of industry and their uses have been increasing and expanding continuously. Colemanite, one of the most common boron minerals, has a monoclinic crystal structure with a chemical formula of 2CaO·3B2O3·5H2O and is used usually in the production of boric acid. The present study concerns and investigation of the dissolution mechanism of colemanite in H2SO4 solution and the effect of acid concentration, the effect of SO4−2 ion on the dissolution process, using H2SO4, HCI+H2SO4 and H2SO4+Na2SO4 solutions. The analysis of the experimental data shows that increasing H3O+ acid concentration increased the dissolution rate, but increasing SO4−2 concentration reduced dissolution rate because of the precipitation of a solid film of CaSO4 and CaSO4·H2O.  相似文献   

12.
The salt rejection by Shirasu porous glass (SPG) membranes having nano-order uniform pores was investigated for understanding the electrokinetic mechanism resulting from the surface charge developed on the membrane when in contact with salt solutions. Due to the dissociation of the hydroxyl groups such as silanol groups on the membrane surface, the membrane was negatively charged over a pH range of 3–10 from electrophoretic measurements. Cross-flow filtration experiments showed that up to 63% of NaCl was rejected by an SPG membrane having a mean pore diameters of 33 nm in a 1 mol m−3 NaCl solution at pH 7 under a transmembrane pressure of 74 kPa, even though the pore diameter is much larger than the ion diameter. This is a consequence of the electrostatic repulsive interaction between the co-ions (Cl ions) and the membrane surface. At the same pH, the rejection factor of NaCl decreased with increasing salt concentration due to an increase in the ionic strength. More negative charge on the membrane surface at higher pH resulted in higher rejection factors of NaCl for a fixed salt concentration. Higher rejection factors of NaCl by SPG membranes with smaller pore sizes for a fixed concentration are due to the higher ratio of the thickness of the electric double layer (Debye length) to the pore radius. The SPG membrane showed a salt rejection sequence: Na2SO4, NaCl and CaCl2 at the same pH. This is because divalent anions (SO42−) are more strongly repelled by the negatively charged membrane, while divalent cations (Ca2+) adsorb specifically onto the membrane surface than monovalent cations (Na+). The salt rejection factor increased with increasing permeate volume flux. Due to the stronger acidity of the membrane materials, SPG membranes had a higher rejection factor and a lower isoelectric point (IEP < 3) than ceramic membranes.  相似文献   

13.
The temperature–concentration dependence of the electrical conductivity of glasses in the NaPO3–NaF system has been investigated. The regularities revealed are interpreted from the standpoint of the structural microinhomogeneity of glasses, which is due to the formation of polar structural units of the Na+[OPOO2/2], Na2 +[O 2POO1/2], Na+[FPOO2/2], and Na+F types. It is shown that the concentration dependence of the electrical conductivity is governed by the ratio between the concentrations of these structural units.  相似文献   

14.
The electrochemical corrosion behaviors of Ni-based superalloy nanocrystalline coating (NC) fabricated by a magnetron sputtering technique have been investigated in comparison with cast alloy in 0.25 M Na2SO4 + 0.05 M H2SO4 and 0.5 M NaCl + 0.05 M H2SO4 solution, respectively. Compared with cast alloy, the NC coating had a little higher passive current density in Na2SO4 acidic solution, while it had superior resistance to pitting corrosion in NaCl acidic solution. The semiconductive type of passive film of the NC coating was p-type in both acidic solutions, while, that of cast alloy changed from p-type in Na2SO4 acidic solution to n-type in NaCl acidic solution. XPS results indicated that Cr2O3 was the main component for the passive films of the NC coating as well as those of the cast alloy. No chloride ion was found in the passive film of NC coating while it was in the passive film of cast alloy. The chloride ions adsorbing on the surface of cast alloy incorporated into the passive film, which induced the formation of n-type oxide film. The nanocrystallization of Ni-base superalloy obviously weakened the adsorption of chloride ions on surface, which decreased the susceptibility of pitting corrosion in acidic solution.  相似文献   

15.
A single crystal of excessively Ni2+-exchanged zeolite Y (FAU, Si/Al = 1.70) was prepared by exchange of |Na71|[Si121Al71O384]-FAU with an aqueous stream 0.05 M Ni(NO3)2 at 293 K and pH 4.9, followed by vacuum evacuation at room temperature and 1.3 × 10?4 Pa. Its crystal structure was determined by single-crystal synchrotron X-ray diffraction techniques in the cubic space group Fd \(\overline{ 3}\) m and was refined to the final error indices R 1/wR 2 = 0.0554/0.1557 for |Ni24.7(NiOH)12.1(Ni2O(OH)2)4.8(Ni4AlO4)1.7Na17.0(H3O)6.9|[Si117Al75O384]-FAU. Crystal has about 53 Ni2+ ions per unit cell, indicating the uptake of excess Ni(OH)2, perhaps as NiOH+ ions. Some dealumination of the framework occurred during Ni2+ exchange. In this structure, Ni2+ ions occupy sites I, I′, II′, II, and III′. The residual Na+ ions are found at sites II′ and II. Due to the low pH of the Ni2+ exchange solution, some H3O+ ions are observed. Nonframework oxygen atoms as oxide and hydroxide ions and orthoaluminate coordinate to some of Ni2+ ions to give NiOH+, Ni2O(OH)2, and Ni4AlO4 3+ groups.  相似文献   

16.
Electrokinetic and rheological properties of Na-bentonite suspensions were investigated in the presence of various electrolyte solutions including LiCl, NaCl, KCl, NH4Cl, NaClO4, CH3COONa, NaNO3, Na2SO4, Na3PO4, CuCl2, MnCl2, CaCl2, BaCl2, NiCl2 and AlCl3. It was found that divalent cations (Cu2+, Mn2+, Ca2+, Ba2+ and Ni2+) and trivalent cation (Al3+) were potential determining cations for the Na-bentonite suspensions. Trivalent cation, Al3+, changed the surface charge of Na-bentonite from negative to positive. The zeta potential measurements showed that monovalent counter-cations and mono-, di- and tri-valent anions were indifferent ions for the Na-bentonite suspensions. The plastic viscosity and the Bingham yield stress values of the Na-bentonite suspensions were also determined in the presence of electrolyte solutions.  相似文献   

17.
In order to protect Ni–Cr alloys from high-temperature corrosion, a new heat-resistant glass-ceramic coating was developed with a glass matrix synthesized on the basis of a composite R x O–Al2O3–SiO2–TiO2 (R–Li, Na+, K+, Mg2+, Ca2+, Ba2+) system. The special features of the formation of crystalline phases in the glasses in heat treatment and the optimum regime for the formation of a glass ceramic structure are described.Translated from Steklo i Keramika, No. 3, pp. 30–32, March, 1996.  相似文献   

18.
《分离科学与技术》2012,47(14):3755-3776
Abstract

In this study, the zeta potential values of vermiculite and expanded vermiculite were measured to determine the effect of pH, clay concentration, and various mono- and multivalent electrolytes including NaCl, KCl, NH4Cl, NaNO3, NaClO4, Na2SO4, Na2CO3, Na3PO4·12H2O, MgCl2·6H2O, CaCl2·2H2O, BaCl2, SrCl2·6H2O, CuCl2·2H2O, CoCl2·6H2O, NiCl2, AlCl3, and CrCl3·6H2O on the electrokinetic properties of vermiculite samples. It was found that generally the measured zeta potential values of expanded vermiculite for the studied systems were slightly more negative than that of vermiculite. The pH profiles of vermiculite and expanded vermiculite at acidic, natural, and basic pH values were obtained to determine the effect of time on the pH values of clay suspensions. The zeta potential measurements showed that the surface charge of clay particles was negative in water. The isoelectric point of vermiculite and expanded vermiculite were determined as pH 2.30 and 2.57, respectively. Divalent cations (Mg2+, Ca2+, Sr2+, and Ba2+), heavy metal ions (Cu2+, Ni2+, and Co2+) and trivalent cations (Al3+ and Cr3+) were potential determining ions for vermiculite and expanded vermiculite particles. Moreover, divalent and trivalent cations caused the change of surface charge from negative to positive. On the other hand, monovalent cations (Na+, K+ and NH4 +), monovalent anions (Cl?, NO3 ?, and ClO4 ?) and multivalent anions (SO4 2?, CO3 2?, and PO4 3?) acted as indifferent ions for these clay particles.  相似文献   

19.
The corrosion behaviour of several austenitic and ferritic stainless steels was studied in the KCl–NaCl–BaCl2 melt (molar ratio 1:1:1) at 600°C in the absence and presence of 0.1 molal sodium salts with different oxyanions, namely, Na2CO3, Na2O2, Na2SO3, Na2SO4, Na3PO4 and Na4P2O7. The corrosion rate, determined from analysis of the melt by atomic absorption, was found to agree well with that determined from anodic polarization and decreased with increasing percentage Cr in the alloy. The presence of the oxyanions led to a decrease in the corrosion rate in the order: P2O 7 4– 4 3– 3 2– 4 2– 2 2– 3 2– which runs parallel to the order of increasing ability of O2– ion donation and indicates that the inhibition process involves the formation of a passivating film on the surface. All stainless steels were found to suffer a significant selective leaching of chromium and among all the oxyanions tested, only CO 3 2– anions suppressed the dechromization in the KCl–NaCl–BaCl2 melt significantly.  相似文献   

20.
Wei Liu  Wei Zhao  Sujuan Zhang 《Desalination》2009,249(3):1288-1293
In this paper, the photocatalytic degradation of trichlorfon, an organophosphorous pesticide, was studied by using TiO2 as a photocatalyst. The effects of various parameters, such as the amount of the photocatalyst, illumination time, reaction temperature, electron acceptors, metal ions, anions, and initial pH value on the photocatalytic degradation of trichlorfon were investigated. The best conditions for the photocatalytic degradation of trichlorfon were obtained. The results show that the optimum amount of the photocatalyst used is 8.0 g L− 1. The photodegradation efficiency of trichlorfon increases with the increase of the illumination time or reaction temperature. The photodegradation efficiency of trichlorfon is increased rapidly by adding a small amount of H2O2, K2S2O8, KBrO3, Fe3+ and Cu2+, however, with the addition of Na+, K+, Mg2+, Ca2+, Zn2+, Co2+ and Ni2+, or with the addition of trace amount of SO42−, Cl, Br, there are no obvious effects on the photocatalytic degradation reactions. Alkaline mediums are favorable for the photocatalytic degradation of trichlorfon. The possible roles of the additives on the reactions and the possible mechanisms of effect were also discussed.  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