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 289 毫秒
1.
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   

2.
The anodic oxidation of dilute Cu(Hg)-pool electrodes (maximum mole fraction of Cu = 0.031) in aqueous solutions of Na3PO4 and NaOH pH range 10.5–12.2, yielded a series of potential-time transients that are found to correspond to the formation of Cu2O, CuO and Cu(OH)2 at the higher pH value, and Cu2O and two soluble species, believed to be CuO2- and HCuO2-, at the lower pH value. No sign of processes involving Hg appeared even at a mole fraction of Cu as low as 10−4. Theoretical calculation based on a model of competition between species of Cu and Hg for the available OH- predicts that the electrode should behave electrochemically as pure Cu down to a Cu activity of ca. 10−12. For both anodic and cathodic processes i-τ−1/2 plots are found linear, proposing the applicability of the law τ = ain, with n = 2.  相似文献   

3.
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   

4.
Two novel 2-substitutd-8-hydroxyquinoline derivatives were designed and synthesized. Their luminance properties and carrier transporting abilities were studied when integrated in four different organic light-emitting device structures. The four devices are structured as following—device A: indium tin oxide glass substrate (ITO)/2-TNATA/NPB/(2-[2-(9-ethyl-9H-carbazol-2-yl)-vinyl]-quinolato zinc (1) and 2-[2-(4-diphenylamino-phenyl)-vinyl]-quinolato zinc (2))/Alq3/LiF/aluminum (Al); device B: ITO/2-TNATA/NPB/1 or 2/LiF/Al; device C: ITO/2-TNATA/1 or 2/Alq3/LiF/Al; device D: ITO/2-TNATA/1 or 2/Alq3/LiF/Al. In the device A, the maximum brightness of compound 1 is 5857 cd m−2 at 11 V with the luminance efficiency 1.84 cd A−1. For 2 it is 6047 cd m−2 at 10 V with an efficiency of 2.22 cd A−1. In the device B, the maximum brightness of 1 is 745 cd m−2 at 11 V and efficiency is 0.32 cd A−1. These values are 1748 cd m−2 at 10 V and 0.42 cd A−1 for 2, respectively. In the device C, the maximum brightness of 1 and 2 are 8729 and 5838 cd m−2 at 10 V, respectively, with an efficiency of 1.99 and 1.71 cd A−1. In the device D, the maximum brightness and efficiency of 1 are 8512 cd m−2 at 13 V and 3.10 cd A−1, and they are 9818 cd m−2 at 13 V and 0.42 cd A−1, for 2. Our results also show that both 1 and 2 are good bipolar and bifunctional molecules with high hole-transporting and luminance properties.  相似文献   

5.
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   

6.
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   

7.
Superplastic alloys and metals possess the ability to undergo large uniform strains prior to failure. Isothermal superplasticity of sheet metal is a phenomenon due to both peculiar process condition and material intrinsic characteristics. The material must have a grain size of less than 10 μm, the forming temperature of around half the absolute melting point and a very low strain rate (in the order of 10−5–10−3 s−1).The instability of superplastic flow under uniaxial stress state has been the subject of different studies. In this paper, under biaxial stress conditions, instability analysis of superplastic PbSn60 alloy using the finite element method is investigated. An original model has been implemented successfully in commercial finite element code in order to predict the imminent failure of material during superplastic forming processes.  相似文献   

8.
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   

9.
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   

10.
Hydrogen permeation in gamma titanium aluminides   总被引:1,自引:0,他引:1  
The permeation of hydrogen in gamma titanium aluminides was studied using the Devanathan–Stachurski (DS) cell. Thin disk-shaped samples of gamma titanium aluminide with three different microstructures were appropriately prepared and cathodically charged in an aqueous 0.1 N NaOH solution. The permeation current was monitored as a function of time. Permeation parameters were calculated using the time lag criterion (non-steady state lag). Values of the apparent diffusion coefficient of hydrogen in gamma titanium aluminide varying from 1.87 × 10−7 cm2/s to 3.75 × 10−6 cm2/s were obtained. This slight variation is attributed to differences in microstructure.  相似文献   

11.
Adsorption of platinum(IV) onto D301R resin   总被引:1,自引:0,他引:1  
Pt(IV) was quantitatively adsorbed by D301R resin in the medium of pH = 3.47. The statically saturated adsorption capacity is 410 mg/g. Pt(IV) adsorbed on D301R resin can be eluted by 1.0-2.0 mol/L NaOH. The rate constant is k298 = 5.43 × 10−5S−1. The adsorption of Pt(IV) on D301R resin obeys the Freundlich isotherm. The adsorption parameters of thermodynamics are as follows: enthalpy change ΔH = 4.37 kJ/mol, Gibbs free energy change ΔG = −5.39 kJ/mol, and entropy change ΔS = 32.76 J/(mol·K). The apparent activation energy is Ea = 22.5 kJ/mol. The coordination molar ratio of the functional group of D301R resin to Pt(IV) is 2:1.  相似文献   

12.
The thermal expansion of U2Fe13.6Si3.4 and Lu2Fe13.6Si3.4 has been measured by X-ray powder diffraction. Both compounds exhibit a large spontaneous magnetostriction. In the ground state, the volume effect 11.2 × 10−3 in U2Fe13.6Si3.4 consists of almost equal contributions from the Fe–Fe and U–Fe exchange interactions (6 × 10−3 and 5 × 10−3, respectively). In Lu2Fe13.6Si3.4, the volume effect is 8.9 × 10−3.  相似文献   

13.
Self-doped polyaniline (SDPA) nanofibers were deposited on platinum (Pt) electrode by reverse pulse voltammetric (RPV) method and their electrochemical performance was evaluated in an aqueous redox supercapacitor constituted as a two electrode cell in a weak acidic medium. Scanning electron micrographs clearly revealed the formation of nanofiber structures with diameters in range of 60–90 nm under optimum experimental conditions. Different electrochemical methods including galvanostatic charge–discharge (CD) experiments, cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) were carried out in order to investigate the applicability of the system as a redox supercapacitor. Based on the charge–discharge results obtained, the SDPA represented high specific capacitance, specific power and specific energy values of 480 F g−1, 436 W kg−1 and 9.40 Wh kg−1, respectively, at a current density of 5 mA cm−2. The present study introduces new nanofiber materials for electrochemical redox capacitors with advantages including low cost, long cycle-life and stable at low acidic solutions of pH 3.  相似文献   

14.
SiC powder prepared by the Na flux method at 1023 K for 24 h and Ba were used as starting materials for synthesis of tribarium tetrasilicide acetylenide, Ba3Si4C2. Single crystals of the compound were obtained by heating the starting materials with Na at 1123 K for 1 h and by cooling to 573 K at a cooling rate of −5.5 K/h. The single crystal X-ray diffraction peaks were indexed with tetragonal cell dimensions of a = 8.7693(4) and c = 12.3885(6) Å, space group I4/mcm (No.140). Ba3Si4C2 has the Ba3Ge4C2 type structure which can be described as a cluster-replacement derivative of perovskite (CaTiO3), and contains isolated anion groups of slightly compressed [Si4]4− tetrahedra and [C2]2− dumbbells. The electrical conductivity measured for a not well-sintered polycrystalline sample was 2.6 × 10−2–7 × 10−3 S cm−1 in the temperature range of 370–600 K and slightly increased with increasing temperature. The Seebeck coefficient showed negative values of around −200 to −300 μV K−1.  相似文献   

15.
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   

16.
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   

17.
Iron molybdenum alloys were prepared for the molybdenum concentration range 0–50 at.% by the arc melting method. X-ray diffraction patterns show single BCC phase for the Mo concentration up to 18 at.% with the lattice constant increasing upon addition of Mo. Sample with 21 at.% of Mo looks like composed of many BCC phases differing by the lattice constant. Finally, sample with 50 at.% of Mo looks like being composed of two BCC phases with various lattice constants and some amount of the non-stoichiometric λ-phase having symmetry P63/mmc. Mössbauer data indicate random solutions up to 12 at.% of Mo with magnetic order at room temperature. Room temperature paramagnetic phase appears for the sample with 18 at.% of Mo and its content increases with the increasing concentration of Mo. Traces of magnetically ordered phase (at room temperature) are seen for the sample with 40 at.% of Mo. Sample with 50 at.% of Mo is paramagnetic at room temperature. Contributions to the hyperfine field and isomer shift on the iron nuclei have been determined as the function of the distance between iron nucleus and Mo impurity up to the third co-ordination shell within the single-phase random solution range. Mo atom as the nearest iron neighbor changes iron hyperfine field by −4.18 T, as the second neighbor makes change by −2.30 T and finally as the third neighbor changes the field by +0.51 T. Corresponding changes in the isomer shift are as follows: −0.033 mm/s, −0.005 mm/s and +0.003 mm/s. The average hyperfine field and the isomer shift decrease linearly versus Mo concentration at rates −0.383 T/at.% and −5.06 × 10−4 mm/(s at.%), respectively. Hence, addition of Mo increases the electron density on the iron nucleus at the rate +1.7 × 10−3 electron a.u.−3(at.%)−1.  相似文献   

18.
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   

19.
The effects of environmental parameters, such as the impurity concentration, oxygen content and pH value on the stress corrosion cracking (SCC) of 3.5Ni-Cr-Mo-V and 2Cr-Ni-Mo steels were studied. SCC of these two alloys only occurs in steams containing both oxygen and contaminating ions. A concentration of 40 ppb of Cl or around 5 ppm of SO42− in steam is found to be enough to induce SCC on both alloys when oxygen exists in the steam. Sulfate ions suppress the severity of chloride ions in causing SCC of both materials in a Cl and SO42− containing steam. This effect was proposed to result from the iron sulfate salt deposition. SCC of these steels in steam is accomplished by corrosion of active path, while the reaction rate is determined by the rate of cathodic reaction of the SCC system. Unless the pH value is 3 or less, the change in pH of de-aerated environments induces no SCC on both materials. SCC in the pH 3 environment is a result of the change of the cathodic reaction from water decomposition to proton reduction in this SCC system.  相似文献   

20.
A rotating disk electrode technique is used to investigate the kinetics and mechanism of O2 reduction as it occurs at the surface of various hot-dip Al–Zn alloy coatings (on steel) immersed in weakly alkaline (pH 9.6) aqueous sodium chloride. The zinc component of coatings behaves electrochemically as though it were free zinc and the O2 reduction pathway is determined by the potential dependent state of zinc. A 2e reduction to H2O2 predominates at potentials near the free corrosion potential, where zinc is (hydr)oxide covered. A 4e reduction to OH predominates at potentials where zinc is bare. Tafel slopes (∂E/∂log i) of 0.058 V dec−1 and 0.132 V dec−1 are determined for 2e and 4e O2 reduction on pure zinc, respectively. Aluminium is virtually inert and varying aluminium content between 0.1% and 55% exerts little influence on O2 reduction kinetics. However, all the Zn–Al alloy surfaces give very much higher O2 reduction currents at low polarization than does pure zinc and it is proposed that this arises through an electrocatalysis of 2e O2 reduction by traces of substrate derived 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