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 312 毫秒
1.
The present paper focuses on the study of SCC behaviour of a new Al–Cu–Li alloy. For this purpose, two conventional media – NaCl and NaCl + H2O2 – were used for comparison with commercial alloys 7075 and 8090. This new alloy shows lower susceptibility to SCC than conventional alloys as it does not undergo environmentally-induced embrittlement in NaCl solutions and in 1 M NaCl + 0.3% H2O2 in which the 7075 and 8090 alloys, respectively, undergo environmentally-induced fracture.Solution composition was modified in order to determine the environmental conditions and strain rates under which this new alloy will crack due to a stress corrosion cracking phenomenon. The addition of 0.6 M sulphates to 1 M NaCl + 0.3% H2O2 solution allows the definition of a range of strain rate (between 10−7 and 10−6 s−1) in which this new alloy undergoes stress corrosion cracking.  相似文献   

2.
Surface resolidification experiments using a high power CO2-laser have been performed on an Al2O3ZrO2 containing 36.8 at.% ZrO2 eutectic alloy at beam velocities between 0.3 and 8 mm·s−1. The local growth rate has been measured by observation of the orientation of the microstructure using scanning electron microscopy. In the whole range of velocities, the structure is essentially a regular lamellar eutectic and the value of the growth productλ2V was found to be ≈ 9.6·10−17 m3·s−1. The measured eutectic spacings were compared with Jackson and Hunt model. Using thermophysical properties from the literature, the measured spacings were more than four times larger than the calculated ones. Assuming all parameters of the growth relationship except the diffusion coefficient to be of the right order of magnitude or to have a negligible influence, agreement is found when using a larger liquid diffusion coefficient,DL≈5·10−10m2·s−1.  相似文献   

3.
Enthalpy increment measurements on CeTe2O6 (s) and ThTe2O6 (s) were carried out using a Calvet micro-calorimeter. The enthalpy increment values were least squares analyzed, with the constraint that H0(T)−H0 (298.15 K) at 298.15 equals 0 and Cp0 (298.15 K) is equal to the value estimated by Kellog’s method.The dependence of the enthalpy increment with temperature can be given as:
H0(T)−H0 (298.15 K)(J mol−1)=189.95 T (K)+15.226×10−3 T2 (K) +15.414×105/T (K)−63157 (CeTe2O6 (s), 391.5–848.0 K)
H0(T)−H0 (298.15 K)(J mol−1)=191.34 T (K)+14.993×10−3 T2 (K) +14.668×105/T (K)−63300 (ThTe2O6 (s), 391.5–909.3 K)
Molar heat capacity Cp0(T), S(T) and Gibb’s free energy functions were evaluated.  相似文献   

4.
New pyrophosphate Sn0.9Sc0.1(P2O7)1−δ was prepared by an aqueous solution method. The structure and conductivity of Sn0.9Sc0.1(P2O7)1−δ have been investigated. XRD analysis indicates that Sn0.9Sc0.1(P2O7)1−δ exhibits a 3 × 3 × 3 super structure. It was found that Sn0.9Sc0.1(P2O7)1−δ prepared by an aqueous method is not conductive. The total conductivity of Sn0.9Sc0.1(P2O7)1−δ in open air is 2.35 × 10−6 and 2.82 × 10−9 S/cm at 900 and 400 °C respectively. In wet air, the total conductivity is about two orders of magnitude higher (8.1 × 10−7 S/cm at 400 °C) than in open air indicating some proton conduction. SnP2O7 and Sn0.92In0.08(P2O7)1−δ prepared by an acidic method were reported fairly conductive but prepared by similar solution methods are not conductive. Therefore, the conductivity of SnP2O7-based materials might be related to the synthetic history. The possible conduction mechanism of SnP2O7-based materials has been discussed in detail.  相似文献   

5.
SYMADEC 1, used previously to synthesise, match and deconstruct polarization curves for the Fe/H2O/H+/O2(4e) corrosion system, has been updated to SYMADEC Multimetal. The polarization curve of a zinc anode alloy corroding with formation of a non-protective layer of ZnO in fully-aerated, near-neutral NaCl solution was modelled. To test the SYMADEC Multimetal model four published polarization curves of zinc actively corroding in similar salt solutions are deconstructed. The polarization curve current/voltage contribution from the ZnO reduction/oxidation peak overlaid on the polarization curve is estimated by a bi-Gaussian relationship.  相似文献   

6.
A Fe---26 Cr alloy has been oxidized at 600°C in 5 × 10−3, 5 × 10−2 and 5 × 10−1 torr oxygen to examine the influence of the prior oxide film on the growth and structure of oxides formed at high temperature. Different prior oxides were produced either by electropolishing or by annealing the electropolished specimen in vacuum at 600°C. Auger electron spectroscopy (AES) showed the Cr content of the prior oxide film to be increased from 50 to 95% during annealing, and electron diffraction indicated a change in oxide structure from amorphous to crystalline. At 5 × 10−3 torr, electropolished Fe---26 Cr oxidizes faster than the vacuum annealed specimens because the amorphous prior oxide gives rise to a finer-grained cubic oxide with more grain boundary easy diffusion paths for cation transport. From AES and electron back-scattering Fe57 Mössbauer spectroscopy it is concluded that this cubic oxide is a duplex layer of inner γ-Cr2O3 and outer Fe3O4. The oxidation rate slows markedly when nucleated α-Fe2O3 covers the cubic oxide. With increased oxidation time Fe3O4 converts to α-Fe2O3 and the γ-Cr2O2 to α-Cr2O3. Annealed Fe---26 Cr oxidizes slower primarily because of a lower cation transport through a coarser-grained cubic oxide rather than because of a higher Cr content in the prior oxide. α-Fe2O3 nucleates at an earlier stage in the oxidation and essentially stifles the reaction. The extent of Cr incorporation into any of the Fe oxides produced in 5 × 10−3 torr oxygen is small ( 5%). Increasing the oxygen pressure from 5 × 10−3 to 5 × 10−2 and 5 × 10−1 torr has little effect on the mechanism of oxidation of vacuum annealed Fe---26 Cr, except that the overall extent of oxidation is less because of earlier α-Fe2O3 formation and, after a few hours of oxidation, up to 20% Cr is incorporated into the α-Fe2O3 lattice. On electropolished Fe---26 Cr at 5 × 10−2 and 5 × 10−1 torr oxygen nodules of α-Cr2O3 form and continue to grow both at grain boundaries and within the grains. Possible mechanisms for this nodule formation, which is exclusive to electropolished specimens oxidized at the higher pressures, are considered.  相似文献   

7.
The effect of partial replacement of copper by zinc in Tl2Ba2Ca2Cu3O10−δ superconductor phase is studied. Superconducting samples of the nominal composition Tl2Ba2Ca2Cu3−xZnxO10−δ with x ranging from 0 to 0.6 are prepared under normal pressure by a one step of solid-state reaction technique. The samples are characterized by using X-ray powder diffraction, scanning electron microscope (SEM) and EDX. The X-ray data indicate that the partial replacement of Cu2+ions by Zn2+ions does not influence the tetragonal structure of the samples, and the lattice parameters a and c vary according to the difference in the ionic radii of Cu and Zn. The superconducting parameters, such as superconducting transition temperature Tc, critical current density Jc and irreversibility field Bir are calculated from electrical resistivity and AC-magnetic susceptibility measurements.  相似文献   

8.
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   

9.
Compositions of surface films formed on nickel and Alloy 600 in I M HCI, 0.5 M H2SO4 and I/3 M H3PO1 solutions were investigated as a function of polarization potential. The main constituent of surface films formed on Ni in 0.5 M H2SO4 or 1/3 M H3PO4 solution was hydrated nickel oxyhydroxide, in which the ratio of O2 to OH increased when passivation occurred. The surface films formed on Ni and Alloy 600 at lower potentials in 0.5 M H2SO4 solution contain S2 ions other than SOP42− ions, whereas S2− ions were not incorporated in the passive film. Passivation of Alloy 600 took place by the formation of hydrated chromium oxyhydroxide. Pitting led to no substantial change in the average composition of the film.  相似文献   

10.
A series of Fe–15Cr–(2–3)Mo–(0.7–2.5)C (compositions in weight percent) steels was oxidised at 850°C and PO_2 = 5.8 × 10–20 atm, where iron oxide is unstable. All grew external Cr2O3 scales according to parabolic kinetics. Depletion of chromium from alloy subsurface regions led to dissolution of chromium-rich carbides if the original alloy carbon level was less than 1.2%. Simultaneous decarburisation caused a transformation of the original austenitic or austenoferritic structure into single-phase ferrite, stabilised by the molybdenum. Diffusion analysis of the concentration profiles within this transformed zone led to satisfactory agreement with the known diffusion coefficient for chromium in ferrite. At high carbon levels, decarburisation was slow, resulting in low chromium concentrations at the internal alloy–carbide interfaces. In these cases, the carbide dissolution did not proceed and chromia scaling rates were slowed.  相似文献   

11.
In this study the different surface states that manifest in the corrosion process of 1018 carbon steel in alkaline sour environment, solution prepared specifically to mimic the sour waters occurring in the catalytic oil refinery plants of the Mexican Oil Company (PEMEX) (0.1 M (NH4)2S and 10 ppm NaCN at pH 9.2) were prepared and characterized. The surface states of the carbon steel were formed by treating the surface with cyclic voltammetry at different switching potentials (Eλ+), commencing at the corrosion potential (Ecorr=−0.890 V vs sulfate saturated electrode, SSE). The surface states thus obtained were characterized using electrochemical impedance spectroscopy and scanning electron microscopy techniques. It was found that for Eλ+=−0.7 and −0.6 V vs SSE a first product of corrosion formed, characterized by a high passivity. Moreover, it was very compact (with a thickness of 0.047 μm). However, at more anodic potentials (Eλ+>−0.5 V vs SSE) a second corrosion product with non-protective properties (porous with a thickness of 0.4 μm and very active) was observed. The diffusion of atomic hydrogen (H0) was identified as the slowest step in the carbon steel corrosion process in the alkaline sour media. The H0 diffusion coefficients in the first and second products that formed at the carbon steel–sour medium interface were of the order of 10−15 and 10−12 cm2/s respectively.  相似文献   

12.
Kunitsugu Aramaki   《Corrosion Science》2006,48(12):4303-4315
A self-assembled monolayer (SAM) of 16-hydroxyhexadecanoate ion was prepared on a zinc electrode coated with a hydrated cerium(III) oxide Ce2O3 layer. The protective efficiency, P of the duplex film composed of the Ce2O3 layer and the SAM was examined by polarization measurement of the zinc electrode in aerated 0.5 M NaCl. The P value of the HOC16A SAM on the Ce2O3-coated electrode was only a little higher than that of the Ce2O3-coated one. The contact angle of the surface covered with the film with a drop of water was far higher than the value of the HOC16A SAM on passivated iron, suggesting that HOC16A was partly adsorbed on the Ce2O3 surface via both and OH groups in a looped orientation. Because this orientation must be changed to an extended orientation during immersion in 0.5 M NaCl, some defects appeared within the SAM in places, resulting in a poor enhancement of P. The film structure was characterized by X-ray photoelectron and FTIR reflection spectroscopies.  相似文献   

13.
Thermoelectric properties of Sn1−xyTiy SbxO2 ceramics were investigated in detail. The addition of Sb into SnO2 matrix increased the electric conductivity, σ. The increase in the σ value should be caused by the increase in the carrier concentration. The Seebeck coefficients of all the samples were negative, which means that these samples have n-type conduction. The samples of this study have porous structure. The maximum Z value of all the samples measured in this study was 2.4 × 10−5 K−1.  相似文献   

14.
The addition of Ni leads to the formation of protective rust layer on steel and subsequently high corrosion resistance of steel in Cl-containing environment. α-FeOOH, β-FeOOH, γ-FeOOH and Fe3O4 are formed mainly on steels exposed to Cl-containing atmosphere. It is expected that systematic investigation of the effect of Ni(II) on the formation process of each oxide in solution should lead to elucidation of the role of Ni in the formation of anticorrosive oxide layer. This study reports the oxidation behavior of NixFe1−x(OH)2 in Cl-containing solution at two different pH regions (condition I under which solution pH is allowed to decrease and condition II under which solution pH is maintained at 8) where γ-FeOOH and Fe3O4 are predominantly formed, respectively, upon the oxidation of Fe(OH)2. In the presence of Ni(II) in the starting solution, the formation of Ni(II) doped β-FeOOH with very low crystalline was facilitated and the formation of γ-FeOOH was suppressed with increasing Ni(II) content and with increasing oxidation rate of Fe(II). Ni(II) was found to have Fe3O4-suppressing effect under condition II.  相似文献   

15.
T. Dudziak  H.L. Du 《Corrosion Science》2009,51(5):1189-1196
In this paper, we present the sulphidation/oxidation behaviour of a Ti45Al8Nb (at%) alloy coated with different protective surface films. Two intermetallic coatings are considered; TiAlCr and Al2Au deposited by physical vapour deposition. The coated alloy was subjected to a H2/H2S/H2O yielding pS2 - 10−1 Pa and pO2 - 10−18 Pa potentials at 750 °C for up to 1000 h. The corrosion kinetics were determined by means of discontinuous gravimetry and the as-received and exposed samples were characterised using scanning electron microscopy (SEM), energy dispersive X-ray analysis (EDX) and X-ray diffraction analysis (XRD). The materials showed the development of a multilayered structure. In the case of the TiAlCr coated Ti45Al8Nb - base alloy, Al2O3, TiO2 and Cr2S3 developed. For the Al2Au coated Ti45Al8Nb samples an Al2O3 scale containing TiO2 nodules was observed at the surface.  相似文献   

16.
Atmospheric corrosion of zinc induced by runoff   总被引:1,自引:0,他引:1  
Atmospheric corrosion and runoff of zinc were investigated during two years in humid tropical climate on hot dip galvanized steel and zinc samples. The high zinc mass loss (14.70 g m−2) is induced by the intensive zinc release (12.40 g m−2). No corrosion phase containing chloride was detected on the zinc surface, while a variety of sulfates not dissolved by rains reveals the sensitivity of zinc to SO2 pollutant. However, two chloride-containing corrosion products were detected on the galvanized steel. Exponential equation is proposed that fits well the experimental data for zinc mass loss induced by runoff process as a function of the time of wetness. The formula gives possibility to predict the mass loss even before a steady state in the corrosion process has been reached. This equation can converge to a Benarie lineal function (C = Atw), when the coefficient b = 1 for the corrosion which is accelerated with the partial removal of the corrosion layer during the runoff phenomena.  相似文献   

17.
The dissolution process of nickel in liquid Pb-free 87.5% Sn–7.5% Bi–3% In–1% Zn–1% Sb and 80% Sn–15% Bi–3% In–1% Zn–1% Sb soldering alloys has been investigated by the rotating disc technique at 250–450 °C. The temperature dependence of the nickel solubility in soldering alloys obeys a relation of the Arrhenius type cs = 4.94 × 102 exp(−39500/RT)% for the former alloy and cs = 4.19 × 102 exp(−40200/RT)% for the latter, where R is in J mol−1 K−1 (8.314 J mol−1 K−1) and T is in K. Whereas the solubility values differ considerably, the dissolution rate constants are rather close for these alloys and fall in the range (1–9) × 10−5 m s−1 at disc rotational speeds of 6.45–82.4 rad s−1. Appropriate diffusion coefficients vary from 0.16 × 10−9 to 2.02 × 10−9 m2 s−1. With both alloys, the Ni3Sn4 intermetallic layer is formed at the interface of nickel and the saturated or undersaturated melt at dipping times of 300–2400 s. The other Ni–Sn intermetallic compounds are found to be missing. A simple mathematical equation is proposed to evaluate the Ni3Sn4 layer thickness in the case of undersaturated melts. The tensile strength of the nickel-to-alloy joints is 94–102 MPa, with the relative elongation being 2.0–2.5%.  相似文献   

18.
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   

19.
C.J. Rosa   《Corrosion Science》1982,22(12):1081-1088
The oxidation of pure Ni has been studied over a temperature range of 1173–1573 K under 1.01 bar oxygen. The oxidation kinetics, the oxide growth and the metal recess obey parabolic time relationships. The dependence of the parabolic rate constant for oxidation can be expressed as: Kp = 7.95 × 10−2 exp (− 220.9/RT) with the activation energy in kJ mole−1. An attempt is made to calculate Kp in terms of diffusion parameters. These calculated values compare well with those measured experimentally.  相似文献   

20.
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   

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