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 15 毫秒
1.
Corrosion cells resulting from differential pH values have been investigated in the absence and presence of Cl ions. Measurements of potential, galvanic current for separate and coupled electrodes, as well as experimental determination of Evans diagrams, were carried out. The results of coupling rvealed that the steel at lower pH (7 and 8) suffered from corrosion while that at the. higher pH (12 and 12·5) was completely protected even in presence of high chloride concentrations The increase in the pH range caused a relative enhancement of the anodic process. The presence of 10−3 or 10−1M Cl ions increased the rate of the anodic reaction by 8- or 12-fold. The rate of the cathodic reaction was of the same order of magnitude both in presence or absence of Cl ions. The % anodic and cathodic control were found to be 6·4 and 93.6, respectively, in the presence of Cl ions as compared with 34·3 and 65·7 in the absence of Cl ions.  相似文献   

2.
The reaction between Fe and HNO3 is studied under a wide variety of conditions by the thermometric technique. Up to 4N HNO3 ΔT varies linearly with the normality of HNO3, while in solutions from 6 to 10N HNO3 it is independent of the concentration. Passivity sets in solutions ≥ 10·8N HNO3. Calculations of the reaction number (R.N.) reveal that the maximum rate of metal dissolution occurs in 7·3N HNO3. Fe dissolution in dilute HNO3 is promoted by additions of NO3 and NO2. That the rate-determining step of the autocatalytic process involves HNO2 is supported by the results of addition of urea to the solution. This additive lowers the maximum measured temperature, without affecting the corresponding time necessary to reach it.Additions of HCI, NaCl, H2SO4 and Na2SO4 to dilute HNO3 reduce the dissolution rate of Fe. The effect produced by the salts exceeds that of the acids.In contrast to its action in dilute solutions, the Clion induces pitting corrosion in concentrated HNO3. The attack starts after an induction period which decreases in length as the concentration of HCI is increased. Concentrated HNO3 can tolerate a certain amount of the aggressive agent before attack starts. The concentration “NHCl” which can be tolerated depends on that of the passivator according to logNHCl=a+blog(NN°)HNO3 where a and b are constants, and No is the least concentration of HNO3 necessary to cause passivity. Pitting corrosion in concentrated HNO3 can be initiated also through NaCl. In one and the same acid solution more of NaCl is needed to cause the attack than of HCI.  相似文献   

3.
The dissolution of copper in 0·36 to 3·6 HCl was studied both in static and in flowing solutions using flow rates between 0·098 and 1·14 ms−1, all in the region of laminar flow.Steady state anode potential current curves, potentiodynamic sweep and potentiostatic pulse techniques and impedance measurements, in the range of 0·02–7 kHz, were used. In both static and flowing solutions dissolution of copper occurs to a monovalent state, as chloro-complexes CuCl2 and CuCl2− with exchange currents for the two reactions of 1·9 × 10−5 A mm−2 and 8 × 10−7 Amm−2 respectively in 1·44 MHCl. The cuprous chloride layer first forms a monolayer and subsequently grows to considerable thicknesses. The double layer capacity is constant at 0·17 ± 0·03 μFmm−2 at potentials below multilayer formation and this is interpreted as implying that there is no specific adsorption of chloride ions prior to formation of the cuprous chloride layer. As the flow rate increases the film becomes thinner so delaying the formation of the film of critical thickness required for passivation.  相似文献   

4.
The kinetics of the dissolution of copper single crystal planes in aerated 0·1N H2SO4 containing various concentrations (10−6−10−2M) of Benzotriazole have been studied. The dissolution rates which were controlled by surface reaction, were a function of the temperature, crystallographic orientation and the concentration of Benzotriazole. The stabilities of the crystal planes were in the order (100) > (110) > (111). At 7·5 × 10−3M Benzotriazole, Cu-Benzotriazole film appeared on the surface, bringing mechanical passivity. Benzotriazole acted as cathodic inhibitor at low concentrations and anodic inhibitor at high concentrations. The corrosion potentials of the crystal planes were in the order (100) > (110) > (111) at all concentrations of benzotriazole.  相似文献   

5.
The enthalpy increments and the standard molar Gibbs energies in the formation of LaFeO3(s) have been measured using a high-temperature Calvet micro-calorimeter and a solid oxide galvanic cell, respectively. The corresponding expression for enthalpy increments is given as:
H°(T)−H°(298.15 K)(J mol−1)(±1.2%)=−36887.27+103.53 T(K)+25.997×10−3T2(K)+11.055×105/T(K).
The heat capacity, the first differential of H°(T)−H°(298.15 K) with respect to temperature, is given as:
Cp,m°(T)(JK−1mol−1)=103.53+51.994×10−3T(K)−11.055×105/T2(K).
From the measured e.m.f. of the cell, (−)Pt/(LaFeO3(s)+La2O3(s)+Fe(s))//CSZ//(Ni(s)+NiO(s))/Pt(+), and the relevant ΔfGm°(T) values from the literature, the ΔfGm°(LaFeO3, s, T) was calculated, and is given as:
ΔfGm°(LaFeO3, s, T)(kJmol−1)(±0.72)=−1319.2+0.2317T(K).
The calculated ΔfHm°(LaFeO3, s, 298.15 K) and S°(298.15 K) values obtained using the second law method are −1334.7 kJ mol−1 and 128.9 J K−1 mol−1, respectively.  相似文献   

6.
Magnetite solubility, as a function of temperature and partial hydrogen pressure, with reference to the typical conditions of the operating fluid of a steam generator of a thermal power plant, has been studied by rigorously solving the problem of equilibria and adopting the scheme proposed by Sweeton and Baes [J. Chem. Thermodynamics2, 479 (1970)]. Stoichiometric calculations have proved that magnetite solubility attains its maximum value, which depends on the characteristics of the electrolytic solution, when the temperature is about 100°C, independently of the type of environment. A rigorous pH calculation was carried out using the method of the characteristic function, which can be applied also to complex systems, and assuming that the effect of the ionic strength may be neglected. The main aim of this study, besides helping power plant chemists to select a proper feedwater conditioning, was to calculate the pH, on a molal basis, of a solution through the best-fit of its exact values, as a function of ammonia concentration inside the inverval 1.0 × 10−8 to 9.0 × 10−3 m with a third-degree logarithmic polynomial. The results, which were obtained in the case of a solution containing NH4OH and H2CO3, demonstrate the validity of this technique which allows the pH of a fairly complex system to be computed accurately. It also allows the correct amount of magnetite dissolution products to be evaluated without considering in detail its chemical equilibria when the solution temperature is above 200°C. This remark was derived from the pH calculations of an ammonia containing solution, which showed its independence of partial hydrogen pressure in the high temperature region, at least as far as the interval 0–1 atm was concerned. The determination of the pH, on a molar basis, of a solution at temperatures of 200, 250, 300 and 350°C, contaminated with sea water so that its acid conductivity was 300μΩ−1cm−1, has been performed. These results have shown that the buffering effectiveness of ammonia is negligible when its concentration falls within the interval 1.0 × 10−6 to 2.0 × 10−5 M, whereas in the range 6.0 × 10−5 to 3.0 × 10−4 M, its effect is quite pronounced.  相似文献   

7.
The behaviours of complexation and dissolution of PbCl2 on the surface of galena were investigated to explore the process of hydro-chemical conversion of galena (PbS) in chloride media. By means of solution chemistry calculation, the production and dissolution of the products PbCl2 were studied. And the passivation of the galena was studied by Tafel curve. The results show that PbCl42− is the main form of PbCl2 presented in the saturated potassium chloride (KCl) solution. The PbCl2 crystal is easy to precipitate when the total concentration of chloride ion ([Cl]T) is equal to 0.92 mol/L, and it is inclined to dissolve when [Cl]T is more than 0.92 mol/L. The chloride complexing reaction rate strongly depends on the Fe3+ion concentration when it is less than 6×10−4mol/L, while passivation occurs on the surface of the electrode when Fe3+ concentration is larger than 6×10−4mol/L. The reaction rate increases obviously when KCl is added, since the activity of Cl increases; thus accelerates the dissolution of PbCl2.  相似文献   

8.
By means of the “interruption kinetic technique”, as applied to oxidation of tungsten under 0.048 bar O2, the oxygen diffusion coefficient in growing WO8−x has been determined for the temperature range of 568–908°C and may be expressed as: D = 6.83 × 10−2 exp (−29,890/RT), with the activation energy given in cal/mole. Calculations are made to show the influence of temperature on the concentration of oxygen vacancies in WO3−x, on their free energy of formation and on the ionic conductivity in WO3−x. From the kinetic data of W oxidation at 800°C, prior to interruption-annealing, values of oxygen diffusion coefficients due to oxygen transport via lattice and short-circuit paths have been calculated as functions of time for growing WO3−x. A simplified oxidation model is used for evaluation of the oxidation rate constant of W at 800°C.  相似文献   

9.
Carbonate-containing green rust 1, GR1(CO32−), is prepared by oxidation of Fe(OH)2 in aqueous solution. Ferrous hydroxide is precipitated from NaOH and FeSO4·7H2O solutions and carbonate ions are added as a Na2CO3 solution. For sufficiently large concentrations of sodium carbonate, SO42− ions do not play any role during the oxidation process and, at the end of the first stage of reaction, Fe(OH)2 oxidizes into GR1(CO32−). In the second stage of reaction, GR1(CO32−) oxidizes into α-FeOOH goethite except when the transformation of ferrous hydroxide is partial, which leads to the formation of magnetite. From the X-ray diffraction analysis of GR1(CO32−), lattice parameters of its hexagonal cell are found to be a = 3.160 ± 0.005 Å and C = 22.45 ± 0.05 Å. From the Mössbauer analysis of the stoichiometric GR1(CO32−), which leads to a Fe2+:Fe3+ ratio of 2:1, the chemical formula is established to be: [Fe4(II)Fe2(III)(OH)12][CO3·2H2O]. The 78 K Mössbauer spectrum of the compound can be fitted with three quadrupole doublets, two Fe2+ doublets d1 and D2 corresponding to isomer shifts (IS) of 1.27 and 1.28 mm s−1 and quadrupole splittings (QS) of 2.93 and 2.67 mm s−1, respectively, and one Fe3+ doublet D3 with an IS of 0.47 mm s−1 and QS of 0.43 mm s−1. These three doublets were already used to fit the Mössbauer spectrum of chloride-containing GR1(Cl) [see J.M.R. Génin et al., Mat. Sci. Forum8, 477 (1986) and J.M.R. Génin et al., Hyp. Int. 29, 1355 (1986)]and therefore are characteristic of GR1 compounds. From the recording of electrode potential E and the pH of the suspension versus time during the oxidation, the standard free enthalpy of formation of stoichiometric GR1(CO32−) is estimated to be ΔG °f = − 966.250 cal mol−1. Knowing the chemical formula and ΔG °f of GR1(CO32−) the Pourbaix diagram of iron in carbonate-containing aqueous solutions is drawn.  相似文献   

10.
The corrosion of the two pure metals and of two alloys containing 15 and 30 wt% Nb has been studied at 600–800°C in H2-H2S-CO2 gas mixtures providing 10−8 atm S2 at all temperatures and 10−24 atm O2 at 600°C and 10−20 atm O2 at 700 and 800°C. The corrosion kinetics were rather complex, being sometimes parabolic and in other cases nearly linear. Under a constant temperature the addition of niobium generally reduced the corrosion rate, except at 700°C when pure cobalt corroded more slowly than the two alloys. The corrosion rates for the same material decreased with an increase in temperature under the same sulfur pressure. Except at 800°C under 10−8 atm S2, which is below the dissociation pressure of cobalt sulfide, the scales presented an outer layer of pure cobalt sulfide and an inner layer of complex composition containing a mixture of double sulfide, niobium oxide and in some cases of unreacted metallic cobalt particles. The addition of niobium was generally beneficial, the effect increasing with its concentration in the alloy, but the corrosion rates of the alloys were still much higher than that of pure niobium, mainly as a result of the lack of formation of a protective layer of niobium sulfide. The corrosion behavior is examined with special reference to the consequences of the low solubility of niobium in cobalt and to the relation between the microstructure of the alloys and the scales.  相似文献   

11.
The rates of dissolution of copper single crystal planes in dilute sulphuric acid containing various concentrations (10−8 to 10−2M) of Cl, Br and I are determined at 30°C. The dissolution rates, which are controlled by transport process in solution, are a function of the temperature, stirring rate, oxygen solubility, crystallographic orientation and the concentration of halide ions. Halide ions acted as anodic inhibitor at low concentrations and cathodic inhibitor at high concentrations. The changes in the dissolution rates have been attributed to specific adsorption of halide ions or precipitation of cuprous halides on the surface of the crystal planes.  相似文献   

12.
The effect of Cl ion on the anodic dissolution of iron in H2SO4 solutions containing low H2S level has been studied by electrochemical polarization curve measurements. The total energy and binding energy of the competitive adsorption for Cl and HS ions have been calculated with CNDO/2 method, as well as the net charge distribution of iron atoms at an anodic potential. The results showed that certain concentration of Cl ion inhibit the anodic reaction of iron accelerated by HS. However, when Cl ion reached saturated adsorption, it began to promote the anodic reaction of iron due to the increased negative charge of iron atoms.  相似文献   

13.
In(Ⅲ) was quantitatively adsorbed by iminodiacetic acid resin (IDAAR) in the medium of pH = 4.52. The statically saturated sorption capacity of IDAAR is 235.5 mg·g^-1. 1.0 mol·L^-1 HCl can be used as an eluant. The elution efficiency is 97.9%. The resin can be regenerated and reused without apparent decrease of sorption capacity. The sorption rate constant is k298 = 1.94 × 10-5 s^-1. The apparent sorption activation energy of IDAAR for In(Ⅲ) is 20.1 kJ·mol^-1. The sorption behavior of IDAAR for In(HI) obeys the Freundlich isotherm. The enthalpy change is AH= 17.2 kJ·mol^-1.  相似文献   

14.
Thin films of magnesia were deposited on various substrates using plasma-assisted liquid injection chemical vapor deposition with volatile Mg(tmhd)2·2H2O (1) (tmhd = 2,2,6,6-tetramethyl-3,5-heptanedione). The precursor complexes, Mg2(tmhd)4·(2), and Mg(tmhd)2·pmdien (3) (pmdien; N,N,N′,N″,N″-pentamethyldiethylenetriamine) were prepared from Mg(tmhd)2·2H2O (1). The temperature dependence equilibrium vapor pressure (pe)T data yielded a straight line when log pe was plotted against reciprocal temperature in the range of 360–475 K, leading to standard enthalpy of vaporization (ΔvapH°) values of 59 ± 1 and 67 ± 2 kJ mol 1 for (2) and (3) respectively. Thin films of magnesium oxide were grown at 773 K using complex (1) on various substrate materials. These films were characterized by X-ray diffraction, scanning electron microscopy and energy dispersive X-ray for their composition and morphology.  相似文献   

15.
Novel cobalt base superalloy and its high-temperature flow behavior   总被引:1,自引:0,他引:1  
A novel cobalt base superalloy containing titanium and aluminum was investigated through metallography, tensile test, and high-temperature isothermal compression deformation. The results show that proper content of titanium and aluminum with can improve the strength and ductility of the cobalt base superalloy. The Co3(Ti,Al) compound with L12 structure precipitates when the novel superalloy is aged at 800℃ for 20 h. The e phase with hcp structure also precipitates when the superalloy is deformed by 28% and then aged at 650℃ for 4 h. The e phase can exist at 800℃. The superalloy has excellent high-temperature mechanical properties. Its maximum flow stress at 850℃ is in the range of 360-475 MPa when the strain rate is between 0.0021 and 2.1 s^-1. The flow stress of the superalloy during high-temperature deformation can be described by the Zener-Hollomon parameter with a hot deformation activation energy of 397 kJ.mol^-1.  相似文献   

16.
Metallorganic chemical vapor deposition (MOCVD) was investigated as a more efficient means to fabricate yttria-stabilized zirconia (YSZ) for thermal barrier coating. The MOCVD precursors were Y(tmhd)3 and Zr(tmhd)4 (tmhd, 2,2,6,6-tetramethyl-3,5-heptanedianato) and delivered via aerosol-assisted liquid delivery (AALD). The maximum YSZ coating rate was 14.2 ± 1.3 μm h−1 at 827 °C yielding a layered coating microstructure. The growth was first-order with temperature below 827 °C with an apparent activation energy of 50.9 ± 4.3 kJ mol−1. Coating efficiency was a maximum of approximately 10% at the highest growth rate. While homogeneous nucleation remained a problem, the deposition of YSZ with only minor carbon content was achieved.  相似文献   

17.
Cordierite body with the formulation of 2.8MgO·1.5Al2O3·5SiO2 was prepared from talc and kaolin as the basic raw materials. Following glass crystallization technique the glass powder was successfully heat treated at 900 °C for 2 h to form a single-phase α-cordierite. The crystal structure of α-cordierite was analysed using X-ray diffraction technique and the Rietveld structural refinement method. Differential thermal analysis (DTA), Fourier-transform infrared (FTIR), field emission scanning electron microscopy (FESEM), coefficient of thermal expansion (CTE) and dielectric properties were also performed. Results show that the materials crystallized as a hexagonal structure with space group of P6/mcc and the room temperature lattice parameters are a = 9.743742 (Å) and c = 9.389365 (Å). FTIR analysis on the glass revealed that only silicate species is the only unit that exists in the glass network. DTA also confirmed that α-cordierite completely formed after 13.5 min of isothermal heating at 900 °C. Coefficient of thermal expansion of synthesized α-cordierite is 2.5 × 10−6 °C−1. The dielectric constant is between 5.0 and 5.5 for 1 MHz and 1.8 GHz, respectively, and the dielectric loss is in the range 10−2. FESEM micrographs revealed that the material is fully densified.  相似文献   

18.
Increased turbine inlet temperature in advanced turbines has promoted the development of thermal barrier coating (TBC) materials with high-temperature capability. In this paper, BaLa2Ti3O10 (BLT) was produced by solid-state reaction of BaCO3, TiO2 and La2O3 at 1500 °C for 48 h. BLT showed phase stability between room temperature and 1400 °C. BLT revealed a linearly increasing thermal expansion coefficient with increasing temperature up to 1200 °C and the coefficients of thermal expansion (CTEs) are in the range of 1 × 10− 5–12.5 × 10− 6 K− 1, which are comparable to those of 7YSZ. BLT coatings with stoichiometric composition were produced by atmospheric plasma spraying. The coating contained segmentation cracks and had a porosity of around 13%. The microhardness for the BLT coating is 3.9–4.5 GPa. The thermo-physical properties of the sprayed coating were investigated. The thermal conductivity at 1200 °C is about 0.7 W/mK, exhibiting a very promising potential in improving the thermal insulation property of TBC. Thermal cycling result showed that the BLT TBC had a lifetime of more than 1100 cycles of about 200 h at 1100 °C. The failure of the coating occurred by cracking at the thermally grown oxide (TGO) layer due to severe oxidation of bond coat. Based on the above merits, BLT could be considered as a promising material for TBC applications.  相似文献   

19.
The ability of a SS 316L surface wetted with a thin electrolyte layer to serve as an effective cathode for an active localized corrosion site was studied computationally. The dependence of the total net cathodic current, Inet, supplied at the repassivation potential Erp (of the anodic crevice) on relevant physical parameters including water layer thickness (WL), chloride concentration ([Cl]) and length of cathode (Lc) were investigated using a three-level, full factorial design. The effects of kinetic parameters including the exchange current density (io,c) and Tafel slope (βc) of oxygen reduction, the anodic passive current density (ip) (on the cathodic surface), and Erp were studied as well using three-level full factorial designs of [Cl] and Lc with a fixed WL of 25 μm. The study found that all the three parameters WL, [Cl] and Lc as well as the interactions of Lc × WL and Lc × [Cl] had significant impact on Inet. A five-factor regression equation was obtained which fits the computation results reasonably well, but demonstrated that interactions are more complicated than can be explained with a simple linear model. Significant effects on Inet were found upon varying either io,c, βc, or Erp, whereas ip in the studied range was found to have little impact. It was observed that Inet asymptotically approached maximum values (Imax) when Lc increased to critical minimum values. Imax can be used to determine the stability of coupled localized corrosion and the critical Lc provides important information for experimental design and corrosion protection.  相似文献   

20.
The properties of a series of lanthanide hexacyanoferrate(III) n-hydrates were studied by means of thermal analysis, IR spectroscopy, Raman spectroscopy and X-ray crystallography. Thermal analyses showed that there were two kinds of complexes in this series, Ln[Fe(CN)6]·5H2O (Ln=La–Nd) and Ln′[Fe(CN)6]·4H2O (Ln′=Sm–Lu). The boundary complex between them was Nd[Fe(CN)6]·5H2O. The IR spectra of the two kinds of complexes were obviously different. For the pentahydrates, there were two sharp CN stretching bands at 2050 and 2140 cm−1, and one band at 1600 cm−1 assigned to the HOH bending. On the other hand, for the tetrahydrates besides the two CN stretching bands at 2050 and 2140 cm−1, a new band was observed at 1940 cm−1, and the HOH bending band split into three bands around 1600 cm−1. From the X-ray crystal analysis, the structure of the boundary complex Nd[Fe(CN)6]·5H2O was determined. It belonged to hexagonal, P63/m, with a=7.467(2) Å, c=13.793(3) Å and Z=2 (R=0.082, Rw=0.126). Neodymium was nine-coordinated in the form of the NdN6(H2O)3 group. The three coordinated water molecules of the 5H2O complex with Nd have a large value for the equivalent isotropic thermal parameter. One of the three water molecules was dissociated easily and the 5H2O complex changed into the stable 4H2O complex with Nd. The crystal of the 4H2O complex is orthorhombic, and belongs to the space group Cmcm as well as the other Ln[Fe(CN)6]·4H2O (Ln=Sm–Lu). Therefore, the structure of Nd[Fe(CN)6]·5H2O is regarded as the boundary structure.  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