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 250 毫秒
1.
The hypothesis that water molecules are involved in the anodic dissolution of iron in acidic sulfate solutions is confirmed experimentally. The oxidation of Fe0 iron to Fe+ is shown to proceed in stages: at the first stage, the interaction between the metal and water molecules results in the formation of surface charge-transfer complexes, which then decompose to form the adsorbed OH groups. The charge transfer in the complexes is estimated to be about 0.5.  相似文献   

2.
A method was developed to characterize and quantify iron corrosion products in clean and sulphide polluted sea water. This method is based upon a selective dissolution with suitable reagents (methanol, glycine, bromine-methanol, hydrochloric acid etc.) of the various compounds and the subsequent chemical analysis of the various dissolved elements. The information thus obtained is integrated by a diffractometric X-ray analysis. The following corrosion products were found:
  • Fe(OH)2 = Fe3O4 ? FeO(OH) ? Fe2O3 + unidentified compound Cr (probably oxychloride) in sea water with pH = 8.1 having a 7 ppm D.O. content.
FeOC1 ? Fe(OH)2 ? Fe3O4 ? FeO.OH in partially deoxygenated sea water (D.O. = 3 ppm) at pH 8.1; Fe0.95S ? Fe(OH)2 in sea water at pH = 7 with 10 ppm of sulphides; Fe0.95S ? FeS ? Fe3.6 · Fe0.9(O · OH · C1) and FeOC1 ? iron oxisulphide ? Fe3O4 ? FeO(OH) ? Fe2O3 · H2O in sea water at pH = 7 and 10 ppm initial sulphides left to oxidate. Since the method of chemical analysis thus developed supplies quantitative information in iron distribution among the various anions and on the various oxidation forms, it is deemed a useful tool for investigation of the corrosion kinetics and mechanism.  相似文献   

3.
Thermodynamic diagrams of Na-S-Fe-H2O system were constructed to analyze the behavior of sulfur and iron in the Bayer process. After digestion, iron mainly exists as Fe3O4 and Fe2O3 in red mud, and partial iron transfers into solution as Fe(OH)3-, HFeO2-, Fe(OH)4- and Fe(OH)42-. The dominant species of sulfur is S2-, followed by SO42-, and then SO32- and S2O32-. The thermodynamic analysis is consistent with the iron and sulfur species distribution in the solution obtained by experiments. When the temperature decreases, sulfur and iron can combine and precipitate. Controlling low potential and reducing temperature are beneficial to removing them from the solution. XRD patterns show that NaFeS2·2H2O, FeS and FeS2 widely appear in red mud and precipitates of pyrite and high-sulfur bauxite digestion solution. Thermodynamic analysis can be utilized to guide the simultaneous removal of sulfur and iron in the Bayer process.  相似文献   

4.
R. Sabot 《Corrosion Science》2007,49(3):1610-1624
The formation of rust can be simulated by oxidation of aqueous suspensions of Fe(OH)2 obtained by mixing solutions of NaOH and a Fe(II) salt. The aim of this study was to investigate the influence of organic species associated with microbially influenced corrosion. The lactate anion, often used as a carbon and electrons source for the development of microorganisms, was chosen as an example. Then, in the first part of the study, Fe(OH)2 was precipitated using iron(II) lactate and NaOH. Its oxidation process involved two stages, as usually observed. The first stage led to a Fe(II-III) intermediate compound, the lactate green rust, . This compound has never been reported yet. Its existence demonstrates that the GR structure is able to incorporate a very wide range of anions, whatever the size and geometry. The second stage corresponded to the oxidation of . It led to ferrihydrite, the most poorly ordered form of iron(III) oxides and oxyhydroxides. In the second part of the study, the formation of rust in seawater was simulated by oxidation of Fe(OH)2 in an aqueous media containing both Cl and anions. The first stage led to the sulphate green rust, , the second stage to lepidocrocite γ-FeOOH. Small amounts of iron(II) lactate were added to the reactants. Lactate ions did not modify the first stage but drastically perturbed the second stage, as ferrihydrite was obtained instead of γ-FeOOH.  相似文献   

5.
A number of investigations on the mechanism of reaction of nickel with SO2 has been summarized. The calculation results of the equilibrium gas composition in homogeneous SO2+O2 mixtures are described over wide ranges of temperatures (500–1100°C) and initial gas compositions. The Ni–O–S phase diagram at 540°C has been compared with data on the stability of interaction products under conditions close to equilibrium. The catalytic activity of NiO has been verified to accelerate the attainment of thermodynamic equilibrium in the SO2–O2–SO3 system. The most effective catalytic activity of NiO occurred at 650–800°C. A monolayer (6 Å) of NiSO4 was detected on the scale surface by ESCA. This surface phase is assumed to be formed either as an activated complex on the NiO catalyst or as the locally stable NiSO4 phase. Both assumptions lead to a possible recognition of the sulfate intermediate mechanism.  相似文献   

6.
The anodic and cathodic behaviour of iron in sulphate containing electrolytes The formation of Fe2(SO4)3 on passive iron at pH = 1 appears probable from a thermodynamical point of view. At high SO42? concentrations the equilibrium system contains but low concentrations of Fe3+, and no Fe2+ ions, a fact showing the relatively elevated stability of the Fe2(SO4)3 layer on passive iron. In slightly acid solution (pH = 4) the passivity of the iron is determined by iron oxide layers. The formation of FeSO4 from metallic iron and sulphate ions is restricted to the transpassive zone (pH 4 to 7), in alkaline solutions even to the active zone. In the pH region 2 to 14 the passive layer on iron has about the same composition in the systems Fe|H2O + SO42? and Fe|H2O.  相似文献   

7.
Conclusions It was found that tempering processes are similar in quenched steel and nitrided iron — in the first stage of tempering (20–180°) the martensite with nitrogen transforms, with formation of metastable F16N2 and temper martensite; in the second stage (180–300°) the retained austenite decomposes and the Fe16N2 Fe4N transformation occurs; in the third stage (300–550°) the number of lattice defects decreases and the Fe4N particles coalesce. After quenching and tempering at 500–600° the alloy consists of a ferrite—nitride mixture of the type of temper sorbite in carbon steel.Kiev Polytechnical Institute. Translated from Metallovedenie i Termicheskaya Obrabotka Metallov, No. 3, pp. 28–30, March, 1974.  相似文献   

8.
Iron(II-III) hydroxysulphate GR(SO42−) was prepared by precipitating a mixture of Fe(II) and Fe(III) sulphate solutions with NaOH, accompanied in most cases by iron(II) hydroxide, spinel iron oxide(s) or goethite. Its [Fe(II)]/[Fe(III)] ratio determined by transmission Mössbauer spectroscopy was 2±0.2, whatever the initial [Fe(II)]/[Fe(III)] ratio in solution. Proportion of Fe(OH)2 increased when the initial [Fe(II)]/[Fe(III)] ratio increased, whereas proportion of α-FeOOH or spinel oxide(s) increased when this ratio decreased. GR(SO42−) is metastable vs. Fe3O4 except in a limited domain around neutral pH. Precipitation from solutions containing both Fe(II) and Fe(III) dissolved species seems to favour GRs formation with respect to stable systems involving iron (oxyhydr)oxides.  相似文献   

9.
Potential–pH diagrams of the iron–oxalic acid and ferric carbide–oxalic acid systems taking into account the equilibrium iron or ferric carbide and ferrous (II) oxalate are plotted. Kinetic investigations of the anodic dissolution of white iron, containing ferritic (99.975% Fe) and cementite (Fe3C) phases, oxalic acid environment evidence that the alloy is passivated due to the formation of a FeC2O4 polylayer.  相似文献   

10.
The influence of Zn-dopant on the precipitation of α-FeOOH in highly alkaline media was monitored by X-ray diffraction (XRD), 57Fe Mössbauer and Fourier transform infrared (FT-IR) spectroscopies and field emission scanning electron microscopy (FE SEM). Acicular and monodisperse α-FeOOH particles were precipitated at a very high pH by adding a tetramethylammonium hydroxide solution to an aqueous solution of FeCl3. The XRD analysis of the samples precipitated in the presence of Zn2+ ions showed the formation of solid solutions of α-(Fe, Zn)OOH up to a concentration ratio r = [Zn]/([Zn] + [Fe]) = 0.0909. ZnFe2O4 was additionally formed in the precipitate for r = 0.1111, whereas the three phases α-FeOOH, α-Fe2O3 and ZnFe2O4 were formed for r = 0.1304. In the corresponding FT-IR spectra, the FeOH and FeO stretching bands were sensitive to the Zn2+ substitution, whereas the FeOH bending bands of α-FeOOH at 892 and 796 cm−1 were almost insensitive. The Mössbauer spectra showed a high sensitivity to the formation of α-(Fe, Zn)OOH solid solutions which were monitored on the basis of a decrease in Bhf values in dependence on Zn-doping. A strictly linear decrease in Bhf for α-FeOOH doped with Zn2+ ions was measured up to r = 0.0291, whereas for r = 0.0476 and higher there was a deviation from linearity. The presence of α-(Fe, Zn)OOH, α-Fe2O3 and ZnFe2O4 phases in the samples was determined quantitatively by Mössbauer spectroscopy. Likewise, Mössbauer spectroscopy did not show any formation of the solid solutions of α-Fe2O3 with Zn2+ ions. FE SEM showed a strong effect of Zn-doping on the elongation of acicular α-FeOOH particles (500–700 nm in length) up to r = 0.1111. For r = 0.1304 the sizes of ZnFe2O4 particles were around 30–50 nm, and those of α-Fe2O3 particles were around 500 nm, whereas a relatively small number of very elongated α-(Fe, Zn)OOH particles was observed. A possible mechanism of the formation of α-(Fe, Zn)OOH, α-Fe2O3 and ZnFe2O4 particles was suggested.  相似文献   

11.
Previous mechanisms for the anodic dissolution of iron have been reviewed, and experimental evidence to distinguish between them is given. Of the various values found in the literature for the anodic Tafel slope, the values of 30 and 40 mV at (30°C) seem to be the most accurate. These two values have been taken as criteria for the possible mechanisms. New reaction pathways have been suggested which also involve the formation of Fe(OH)2, FeOH+, and Fe2OH+ as intermediates, and both the Tafel slope and the reaction order with respect to OH? for these pathways have been deduced from steady state kinetics. The data thus obtained show that many of the proposed rate-determining steps give rise to the same results derived for previous mechanisms, thus indicating that such previous mechanisms are not the only possible ones. Difficulties associated with the work on iron, and with the present state of steady state kinetics are also outlined.  相似文献   

12.
It has been shown that the dissolution rate for iron at cathodic potential is higher than expected on the grounds of the Wagner-Traud mixed potential theory. This observation cannot be explained by a parallel chemical process as suggested by Kolotyrkin et al. The dissolution appeared to be unaffected by potential, pH, cations (Li+, Na+, K+), anions (SO2?4, Cl?, ClO?4) and solvent (water-dioxan mixtures). However, the dissolution rate was decreased by an order of magnitude when the electrode was connected to a strong magnet, suggesting that the mechanical degradation of the surface under the influence of the cathodically evolved hydrogen is the main cause of the measured effects.  相似文献   

13.
Järdnäs  A.  Svensson  J.-E.  Johansson  L.-G. 《Oxidation of Metals》2003,60(5-6):427-445
The oxidation of Fe was investigated at 500–700°C in the presence of O2 with 0–1000 ppm SO2. The exposures were carried out in a thermobalance and lasted for 24 h. The oxidized samples were investigated by grazing-angle XRD, SEM/EDX, GDOES and XPS. The rate of oxidation of pure iron is slowed down by traces of O2 in O2 below 600°C while SO2 has no effect on oxidation rate at higher temperatures. Exposure to SO2<600°C resulted in the formation of small amounts of sulfate at the gas/oxide interface. In addition, sulfur, probably sulfide, accumulated at the metal/oxide interface. The influence of SO2 on oxidation rate is attributed to surface sulfate. The sulfur distribution in the scale is rationalized in terms of the thermodynamic stability of compounds in the Fe–O–S system. Exposure to SO2 caused the formation of hematite whiskers.  相似文献   

14.
It has been shown that the dissolution rate for iron at cathodic potential is higher than expected on the grounds of the Wagner-Traud mixed potential theory. This observation cannot be explained by a parallel chemical process as suggested by Kolotyrkin et al. The dissolution appeared to be unaffected by potential, pH, cations (Li+, Na+, K+), anions (SO2-4, Cl?, ClO-4) and solvent (water-dioxan mixtures). However, the dissolution rate was decreased by an order of magnitude when the electrode was connected to a strong magnet, suggesting that the mechanical degradation of the surface under the influence of the cathodically evolved hydrogen is the main cause of the measured effects.  相似文献   

15.
Composites with ferromagnetic nanoparticles, Fe and Fe50Ni50, dispersed in Al2O3 have been synthesized by a solution phase technique. The structure and magnetic properties of these composites with varying fractions of Al2O3 have been investigated. Both Fe and Fe50Ni50 nanoparticles are amorphous in the as-prepared state and become crystalline on heat treating with near equilibrium lattice parameters of 0.287 nm and 0.358 nm respectively. The interparticle distance increases with increasing Al2O3 from 0 wt.% to 20 wt.%. The size of Fe nanoparticles is 40 nm while the Fe50Ni50 nanoparticles are 20 nm in size. The Fe and Fe50Ni50 nanoparticles dispersed composites are found to be ferromagnetic at room temperature both in the as-prepared and heat treated conditions with clear coercive fields of 5.5–35 × 103 A m−1. The saturation magnetization increases by orders of magnitude on heat treatment, for e.g. from <1.0 emu g−1 to 143.4 emu g−1 for Fe–15 wt.% Al2O3 and 95.6 emu g−1 for Fe50Ni50–15 wt.% Al2O3. The Fe-composites exhibit a Curie transition at 1000 K while the Fe50Ni50 composites exhibit a transition at 880 K, both temperatures close to bulk values.  相似文献   

16.
The high-temperature-corrosion behavior of alloy 800H has been studied in an oxidizing (SO2–O2, =0.23 atm, =1.9×10–29 atm) and a reducing (H2–H2S–CO–CO2–N2, =1.5×10–18 atm =4.3×10–8 atm, ac=0.03) sulfidizing environment, at 750°C and 850°C, respectively. When corroded in SO2–O2, the protective chromia scale which developed on the alloy in the early stages cracked and spalled in quite a short time period. This led to the growth of iron and nickel sulfides beneath the chromia layer, causing more chromia spallation. When correded in H2–H2S–CO–CO2–N2, the alloy exhibited breakaway corrosion in about 35hr, at which stage liquid nodules formed on the sample surface. The nodules were studied in detail and were found to consist of three layers. The growth mechanism of such nodules is proposed.  相似文献   

17.
The sulfidation of 310 stainless steel was studied over the temperature range from 910 to 1285° K. By adjusting the ratio of hydrogen to hydrogen sulfide, variations in sulfur potential were obtained. The effect of temperature on sulfidation was determined at three different sulfur potentials: 39 N·m–2, 1.4×10–2 N·m–2, and 1.5×10–4 N·m–2. All sulfide scales contained one or two surface layers in addition to a subscale. The second outer layer (OL-II), furthest from the alloy, contained primarily Fe-Ni-S. The first outer layer0 (OL-I), nearest the subscale, contained Fe-Cr-S. The subscale consisted of sulfide inclusions in the metal matrix. Two different phases were observed in OL-II depending on the temperature and sulfur potential. Below 1065°K OL-II is composed of a mixture of monosulfides of iron and nickel (Fe Ni)1–xS and pentlandite (Fe4.5Ni4.5S8) with the pentlandite phase exsolved as lamellae upon cooling. At temperatures higher than 1065°K only the pentlandite phase was formed, which melted above 1145°K at sulfur potentials greater than 10–2 N·m–2, yielding metal-rich iron-nickel-sulfur. Above 1145°K, and at sulfur potentials less than 10–2 N·m–2, OL-II ceased to exist (this temperature is termed transition temperature). Below the transition temperature, where OL-II exists, OL-I could be represented by the general composition (Fe, Cr)1–xS. This phase on cooling transformed into an array of structures differing in FeCr ratio. These substructures, however, were not observed in quenched samples. Above the transition temperature OL-I changed to a chromium-rich sulfide composition and was associated with a sudden decrease in reaction rate. Subscale formation was found to be due to the dissociation of OL-I at the scale-metal interface, and the extent of subscale growth was found to depend on the temperature and the sulfur potential, as well as the composition of OL-I. At a given temperature and sulfur potential the weight-gain data obeyed the parabolic rate law after an initial transient period. The parabolic rate constants obtained at the sulfur potential of 39 N·m–2 did not show a break when the logarithm of the rate constant was plotted as a function of the inverse of absolute temperature. Sulfidation carried out at a sulfur potential below 2 × 10–2 N·m–2, however, did show a break at 1145°K. This break was found to be associated with the changes which had occurred in the FeCr ratio of OL-I. Below the transition temperature the activation energy was found to be approximately 125 kJ · mole–1. Above the transition temperature the rate of sulfidation decreased with temperature but depended on the FeCr ratio in the ironchromium-sulfide layers of the OL-I. A reaction mechanism consistent with the experimental results has been proposed in which the diffusion of cations through OL-I is the rate-controlling step. Below the transition temperature the diffusion of Fe and Ni through OL-I contributes to the scale formation, whereas above the transition temperature the diffusion of Cr through OL-I controls the scale formation. Existing literature on the Fe-Ni-S system is compared with the present results.  相似文献   

18.
It is shown that, in contrast to aqueous electrolytes, where the electrode process is promoted by the electroreduction of activated surface complexes of Cu2+ with SO 4 2– anions, in aqueous-ethanolic electrolytes, it is promoted by sulfate complexes of copper(II) formed in the solution bulk and needing dissociation prior to the discharge. In the presence of cyclic polyester, the discharging complex species have to permeate through the barrier layer at the cathode surface. The layer consists of ethanol molecules and electroneutral associates of sulfate ions with Cu2+ cations containing crown-ester molecules in the coordination sphere. As a result, the electrode reaction is considerably inhibited with an increase in the concentration of SO 4 2– anions (anionic effect), and tribotechnical characteristics of copper coatings are significantly improved.Translated from Zashchita Metallov, Vol. 41, No. 2, 2005, pp. 162–167.Original Russian Text Copyright © 2005 by Kuznetsov, Skibina, Geshel, Loskutnikova, Kuznetsova.  相似文献   

19.
The corrosion behavior of five Fe-Al binary alloys containing up to 40 at. % Al was studied over the temperature range of 700–900°C in a H2/H2S/H2O mixture with varying sulfur partial pressures of 10–7–10–5 atm. and oxygen partial pressures of 10–24–10–2° atm. The corrosion kinetics followed the parabolic rate law in all cases, regardless of temperature and alloy composition. The parabolic rate constants decreased with increasing Al content. The scales formed on Fe-5 and –10 at.% Al were duplex, consisting of an outer layer of iron sulfide (FeS or Fe1–xS) and an inner complex scale of FeAl2S4 and FeS. Alloys having intermediate Al contents (Fe-18 and –28 at.% Al) formed scales that consisted of mostly iron sulfide and Al2O3 as well as minor a amount of FeAl2S4. The amount of Al2O3 increased with increasing Al content. The Fe 40 at.% Al formed only Al2O3 at 700°C, while most Al2O3 and some FeS were detected at T800°C. The formation of Al2O3 was responsible for the reduction of the corrosion rates.  相似文献   

20.
Anodic and cathodic potentiostatic current transients of freshly formed iron (99.95%) electrodes in 0.5 M aqueous acidic (pH 1.7–3.2) sulfate solutions are examined. The initial stages of iron passivation are confirmed to be determined by the following several processes each prevailing in a definite period: the formation of surface charge-transfer complexes (H2O) δ+ ads (first 3–5 ms) and their transformation into adsorbed OH groups (subsequent 10–50 ms) with the concurrent adsorption of hydrogen atoms Hads by iron.  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