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.
When 20 pct cold-worked Type 316 stainless steel is exposed to Cs at 700°C under controlled oxygen-chemical potential environment, Cs penetration into the stainless steel grain boundaries occurs at oxygen potentials ΔGo2 -96 kcal per mole. At lower oxygen potentials (~ΔGo2 ≤ —110 kcal per mole), no corrosion occurs. Under the same experimental conditions, when the stainless steel is exposed to Cs:Te (2:1, atomic), corrosion occurs and penetration morphology appears to depend strongly on the oxygen-potential environment. The stainless steel suffers intergranular corrosion by Te (in the presence of Cs-Te) under conditions where chromium oxidation is not expected to occur. The kinetics of grain-boundary penetration by Te have been studied at temperatures between 550 and 700°C. The depth of the penetrated zone varies as (time)1/2, and the process has an activation energy of 34 kcal per mole. The results are discussed, and the effects of stainless steel microstructure and externally applied stress on corrosion reactions are also described.  相似文献   

2.
When austenitic stainless steel (300 series) is exposed to cesium oxides in the tempera-ture range from 450° to 700°C the grain boundaries are attacked preferentially. The penetration of cesium oxides into the grain boundaries of AISI Type 304 stainless steel has been studied as a function of time and temperature. These investigations have established that the penetration kinetics are linear in time, the activation energy for the process is 19 kcal/mole, and the rate of penetration is fairly insensitive to carbide precipitation and precipitate composition and morphology. The kinetics of the process are approximately an order of magnitude faster than those observed for some reactor (U, Pu) oxide fuel elements clad with Type 304 stainless steel under fast-flux irradiations, and the results are discussed qualitatively.  相似文献   

3.
The CO(g) pressure in equilibrium with a Ta2C-Ta2O5-Ta mixture has been measured at temperatures between 1740 and 1900 K using the torsion-effusion technique. From the equilibrium data, the following equation for ΔG°2 of Ta2C has been obtained: ΔG°2 (±300) = −47,000 (±2200) +.IT From the enthalpy term in the ΔG°f equation, a value of —47.9 (±2.3) kcal/mole has been calculated for ΔH°298 of Ta2C which is in good agreement with several calorimetric results. This paper is based upon a thesis submitted by A. D. KULKARNI in partial fulfillment of the requirements of the degree of Doctor of Philosophy at the University of Pennsylvania.  相似文献   

4.
The pressures of carbon monoxide in equilibrium with a Cr23C6-Cr2O3-Cr mixture and with a Cr7C3-Cr2O3-Cr23C6 mixture have been measured in the temperature range 1100 to 1300 K using the torsion-effusion technique. From the equilibrium data, the following equation for ΔGof of Cr23C6 (in cal per mole) has been calculated: ΔG f ° (±1200) = −77,000 - 18.3T (1150 to 1300 K) Combining the results of this study at temperatures between 1100 and 1300 K with those of Kelleyet al., 3 at temperatures between 1500 and 1720 K, the following equation for ΔGof of Cr7C3 (in cal per mole) has been determined: ΔG f ° (±400) = −35,200 - 8.7T (1100 to 1720 K) ) The above equation for ΔGof of Cr7C3 has been used to re-evaluate the equilibrium data of Kelleyet al., 3 and the following equation for ΔGof of Cr3C2 (in cal per mole) has been obtained: ΔG f ° (±400) = −16,400 - 4.4T (1300 to 1500 K) CHROMIUM reacts with carbon to form three carbides:1,2 Cr23C6, Cr7C3, and Cr3C2. The chromium carbides are of considerable technical importance because of their precipitation behavior in certain high-chromium steels and superalloys. A precise knowledge of their thermodynamic properties is essential for the understanding and the prediction of their chemical behavior in various environments. This paper is based upon a thesis submitted by A. D. KULKARNI in partial fulfillment of the requirements of the degree of Doctor of Philosophy at the University of Pennsylvania.  相似文献   

5.
The vapor pressures of Na above stirred Na2O-SiO2 melts in equilibrium with graphite and CO were determined at 1300° and 1400 °C using the transpiration technique. Compositions studied ranged from about 60 mole pct SiO2 to close to SiO2 saturation. Activities of components Na2O and SiO2 were calculated from the data. Log aNa2O (pure liquid as standard state) varies from about −8.7 and −8.5 at silica saturation to −6.3 and −6.1 at 40 mole pct Na2O at 1300° and 1400 °C, and the molar Gibbs energy of mixing, ΔG m, at the disilicate composition (XNa2O = 0.33) at each of these temperatures is −83.0 and −85.4 kJ, respectively. The Toop and Samis, Yokokawa and Niwa, and Lin and Pelton solution models for binary silicates were applied to the ΔG m data at 1350 °C and parameters for the models were estimated to give best fits. All three models show good correspondence with the measured ΔG m curve. The capabilities of the models in predicting activity data in this system have been compared. D. N. Rego, Formerly Graduate Student at Carnegie-Mellon University, G.K. Sigworth, Formerly with Carnegie-Mellon University,  相似文献   

6.
A kinetics study of the reduction of Nb2O5 with NH3 was conducted at 600° to 1300°C, using vertical fixed-bed, flow-through reactors, with the goal of using the nitride as an interme-diate in the preparation of niobium (columbium) metal via a thermal decomposition step. The effects of reactor materials (stainless steel, nickel, molybdenum, graphite, alumina, and Vycor) upon ammonia reactivity toward Nb2O5 were investigated. At low temperatures, the metal reactor systems were more catalytically reactive, yielding faster rates of reac-tion and a greater degree of nitride conversion, whereas at high temperatures, the non-metal reactor systems performed better. In general, the initial reaction rate-temperature data exhibited a maximum, associated with oxynitride formation, near 700°C for the metal reactor systems and 800° to 900°C for the nonmetal reactor systems, followed by a mini-mum, associated with NbO2 formation, at 800° to 850°C for the metal reactor systems and 950° to 1000°C for the nonmetal reactor systems where NbN formation commences. A sec-ond maximum, associated with the hexagonal NbN phase, occurred at 1200°C. The ranges of activation energies for these regions were from 15 to 30 kcal/mole for region I, 8 to 22 kcal/mole for region II, and 10 to 22 kcal/mole for region III.  相似文献   

7.
High temperature thermodynamic data for equilibria in the Ca-S-O, Mg-S-O, and La-S-0 systems were determined by a galvanic cell technique using calcia stabilized zirconia (CSZ) solid electrolytes. The measured emf data were used to calculate the standard free energy changes of the following reactions: [1] CaO(s) + 1/2S2(g) → CaS(s) + 1/2O2(g) 1000 to 1350 K ΔG° = 21906.9 − 0.8T(K)(±400 cal) = 91658 − 3.37 (±1700 J) [2] CaS(s) + 2O2(g) → CaSO4(s) 1050 to 1450 K ΔG° = -227530.7 + 80.632T(K) (±400 cal) = -951988.5 + 337.4T (±1700 J) [3] CaO(s) + 3/2O2(g) + 1/2S2(g) → CaSO4(s) 1050 to 1340 K ΔG° = -204892.7 + 79.83T(K)(±400 cal) = -857271.1 + 334.0T (±1700 J) [4] MgO(s) + 1/2S2(g) → MgS(s) + 5O2(g) 1000 to 1150 K ΔG° = 45708.6 − 2.897(K)(±500 cal) = 191244.8 − 12.1T (±2100 J) [5] La2O3(s) + 1/2S2(g) → La2O2S(s) + 1/2O2(g) 1080 to 1350 K ΔG° = 17507 − 2.32T(K)(±380 cal) = 73249.3 − 9.7T (±1600 J) [6] La2O3S(s) + S2(g) → La2S3(s) + O2(g) 950 to 1120 K ΔG° = 70940 + 2.25T(K)(±500 cal) = 296812.9 + 9.47 (±2100 J) The ΔG° values of reaction [5] were combined with the literature data for ΔG°f(La2O3) to obtain the standard free energy of formation of La2O2S at high temperatures. The values of ΔG°f thus calculated for La2O2S were combined with the ΔG° data for reaction [6] to obtain the standard free energy of formation of La2S3 at high temperatures.  相似文献   

8.
The standard Gibbs energies of formation of platinum-rich intermetallic compounds in the systems Pt-Mg, Pt-Ca, and Pt-Ba have been measured in the temperature range of 950 to 1200 K using solid-state galvanic cells based on MgF2, CaF2, and BaF2 as solid electrolytes. The results are summarized by the following equations: ΔG° (MgPt7) = −256,100 + 16.5T (±2000) J/mol ΔG° (MgPt3) = −217,400 + 10.7T (±2000) J/mol ΔG° (CaPt5) = −297,500 + 13.0T (±5000) J/mol ΔG° (Ca2Pt7) = −551,800 + 22.3T (±5000) J/mol ΔG° (CaPt2) = −245,400 + 9.3T (±5000) J/mol ΔG° (BaPt5) = −238,700 + 8.1T (±4000) J/mol ΔG° (BaPt2) = −197,300 + 4.0T (±4000) J/mol where solid platinum and liquid alkaline earth metals are selected as the standard states. The relatively large error estimates reflect the uncertainties in the auxiliary thermodynamic data used in the calculation. Because of the strong interaction between platinum and alkaline earth metals, it is possible to reduce oxides of Group ILA metals by hydrogen at high temperature in the presence of platinum. The alkaline earth metals can be recovered from the resulting intermetallic compounds by distillation, regenerating platinum for recycling. The platinum-slag-gas equilibration technique for the study of the activities of FeO, MnO, or Cr2O3 in slags containing MgO, CaO, or BaO is feasible provided oxygen partial pressure in the gas is maintained above that corresponding to the coexistence of Fe and “FeO.” Formerly Professor and Chairman, Department of Metallurgy, Indian Institute of Science Formerly Visiting Scientist, Department of Metallurgy, Indian Institute of Science  相似文献   

9.
Phase relations and thermodynamic properties of the Cr-O system were studied at temperatures from 1500 °C to 1825 °C. In addition to Cr and Cr2O2, a third crystalline phase was found to be stable in the temperature range from 1650 °C to 1705 °C. The atomic ratio of oxygen to chromium of this phase, which decomposes upon cooling to form Cr and Cr2O3, was determined as 1.33 + 0.02, in good agreement with the formula Cr3O4. Temperatures and phase assem blages for invariant equilibria of the Cr-O system were determined as follows: Cr2O3 + Cr + Cr3O4, 1650 °C ± 2 °C; Cr3O4 + Cr + liquid oxide, 1665 °C ± 2 °C; and Cr3O4 + Cr2O3 + liquid oxide, 1705 °C ± 3 °C. The composition of the liquid oxide phase at the eutectic temperature of 1665 °C was found to be close to CrO. Relations between oxygen pressure and temperature for the univariant equilibria of the Cr-O system were established by equilibrating Cr and/or Cr2O3 starting materials in H2-CO2 mixtures of known oxygen potentials at temper atures from 1500 ΔC to 1825 °C. From this information, the standard free-energy changes (ΔGΔ) for various reactions were calculated as follows: 2Cr (s) + 3/2O2 = Cr2O3 (s): ΔG ° = -1,092,442 + 237.94T Joules, 1773 to 1923 K; 3Cr (s) + 2O2 = Cr2O4 (s): ΔG ° =-1,355,198 + 264.64T Joules, 1923 to 1938 K; and Cr (s) + l/2O2 = CrO (1): ΔG ° =-334,218 + 63.81T Joules, 1938 to 2023 K. Formerly Graduate Research Assistant, The Pennsylvania State University Formerly Professor  相似文献   

10.
The microstructure and phase stability of the Fe-15Mn-7Si-9Cr-5Ni stainless steel shape memory alloy in the temperature range of 600 °C to 1200 °C was investigated using optical and transmission electron microscopy, X-ray diffractometry (XRD), differential scanning calorimetry (DSC), and chemical analysis techniques. The microstructural studies show that an austenite single-phase field exists in the temperature range of 1000 °C to 1100 °C, above 1100 °C, there exists a three-phase field consisting of austenite, δ-ferrite, and the (Fe,Mn)3Si intermetallic phase; within the temperature range of 700 °C to 1000 °C, a two-phase field consisting of austenite and the Fe5Ni3Si2 type intermetallic phase exists; and below 700 °C, there exists a single austenite phase field. Apart from these equilibrium phases, the austenite grains show the presence of athermal ɛ martensite. The athermal α′ martensite has also been observed for the first time in these stainless steel shape memory alloys and is produced through the γ-ɛ-α′ transformation sequence.  相似文献   

11.
Deformation accelerates the development of grain boundary chromium depletion (GBCD), or sensitization, in type 316 austenitic stainless steels (SS). Quantitative assessment of the degree of sensitization (DOS) using the electrochemical potentiokinetic reactivation (EPR) test indicates that the acceleration in GBCD is a function of the amount of strain in the material and temperature of isothermal sensitization treatment. A systematic increase in strain from 0 to 20 pct yields a continuous increase in EPRDOS values below 700°C, while at higher temperatures, a threshold strain of 6 to 10 pct is required to cause accelerated GBCD development. Straining SS above 20 pct also produces higher amounts of chromium depletion, though the (intergranular) sensitization susceptibility of the material could not be quantitatively evaluated due to the presence of grain matrix or transgranular corrosion. Classical C-curve precipitation-sensitization behavior was also noted for strained and unstrained materials, though strain moved the C-curves to the left. Microstructural evaluation of sensitization revealed a systematic increase in grain boundary and twin boundary corrosion on EPR attack surfaces with strain, which corroborated the deformation-induced acceleration of EPRDOS. A time-temperature-strain dependence of transgranular corrosion was also identified on EPR-etched samples strained above 20 pct. These were also reflected in transmission electron microscope (TEM) observations of higher grain boundary carbide precipitation on strainedvs unstrained specimens and site-specific carbide precipitation on deformation sites in the material. Kinetic and thermodynamic modeling of deformation effects on carbide precipitation and depletion development in type 316 SS indicated that strain induces a reduction in the activation barrier to diffusion (Q) a and thermodynamic barrier to nucleation (ΔG *) during the precipitation-depletion process. The lowering ofQ a with strain caused chromium diffusivity and depletion development to be accelerated in strainedvs unstrained materials and appears to be due to increased dislocation pipe diffusion with strain. Reduction of ΔG * with strain was related to an increase in the free energy change of the grain boundary (ΔG) gb and accelerated carbide precipitate nucleation in deformed SS. The effect of strain on the kinetics and thermodynamics of the precipitation-depletion process decreases with increasing temperature.  相似文献   

12.
The nature of the cement copper deposit was shown to control the kinetic response of the cementation reaction under certain circumstances. In essence, the nature of the surface deposit determines the effective cathodic area and is controlled by a number of variables. What appeared to be a temperature region (0 to 35°C) where a surface reaction mechanism was rate controlling, ΔE a ≊10 kcal per mole, was, in fact, a consequence of the variation of the area of the surface deposit with temperature. Further evidence of this phenomenon was demonstrated by the results of cementation experiments in an ultrasonic field, and by the results obtained from initial “strike” experiments. Also considered in this study was the effect of the initial cupric ion concentration and the back reaction kinetics. The basic conclusion reached from this investigation was that the cementation reaction rate is controlled by boundary layer diffusion processes at all temperatures and concentrations with an activation energy of approximately 5 kcal per mole. When surface deposit effects are neglected, interpretation of cementation rate data, as well as rate data of any heterogeneous reaction involving a solid phase, can often be misleading.  相似文献   

13.
It is hypothesized that for ductile austenitic stainless steels exposed to boiling MgCl2 solution, the relevant crack propagation mechanism is slip dissolution. This model relates crack advance to oxidation or anodic dissolution that occurs on the bare surface that is created when a thermo-dynamically stable, protective film at the crack tip mechanically ruptured. Based on the model of slip-bare metal dissolution repassivation and crack-tip strain analysis, a theoretical equation of stress-corrosion crack growth rate as a function of crack-tip strain rate and potential for 321 stainless steel in boiling 42 pct MgCl2 solution is proposed. The theoretical prediction shows that when the crack-tip strain rate changes from 10−4 to 10−2 s−1 the crack propagation rate changes from 0.01 to 3 mm/h at the free corrosion potential (−0.35 VSCE). If the crack-tip strain rate is above 10−2/s, the crack propagation rate should correspond to the upper bound determined by the maximum metal dissolution rate. When the crack-tip rate is below 10−4/s, the crack propagation rate is below 0.01 mm/h. The slip-step dissolution model predicted that there exists a critical potentialE c, above which the crack propagation rate is independent on potential, but below which the crack propagation rate decreased with decreasing potential. The theoretical prediction has been verified by slow strain rate tests of 321 stainless steel under potential control (above −0.35 VSCE) in 42 pct MgCl2 solution.  相似文献   

14.
The standard free energies of formation of calcium phosphide and calcium stannide were determined by a chemical equilibration technique, yielding the following results: 3Ca(1) + P2(g) = Ca3P2(s) ΔG° = −653,460(±7110) + 144.01(±4.98)T (J/mol)1000 °C to 1300 °C2Ca(1) + Sn(1) = Ca2Sn(s) ΔG° = −353,970(±1670) + 79.28(±1.26)T (J/mol)1000 °C to 1300 °C 1120 °C The experimental data to express the thermodynamics for removal of phosphorus and tin from molten iron by calcium based slags by other investigators were discussed in terms of the activity co-efficients of Ca3P2 and Ca2Sn in slag melts by using the present results described above.  相似文献   

15.
The standard free energies of formation of calcium phosphide and calcium stannide were determined by a chemical equilibration technique, yielding the following results: 3Ca(1) + P2(g) = Ca3P2(s) ΔG° = −653,460(±7110) + 144.01(±4.98)T (J/mol)1000 °C to 1300 °C2Ca(1) + Sn(1) = Ca2Sn(s) ΔG° = −353,970(±1670) + 79.28(±1.26)T (J/mol)1000 °C to 1300 °C 1120 °C The experimental data to express the thermodynamics for removal of phosphorus and tin from molten iron by calcium based slags by other investigators were discussed in terms of the activity co-efficients of Ca3P2 and Ca2Sn in slag melts by using the present results described above.  相似文献   

16.
Interaction between molten salts of the type LiCl-KCl-MeCl (Me = Na, Rb, Cs, x MeCl = 0 to 0.5, x KCl/x LiCl = 0.69) and zeolite 4A have been studied at 823 K. The main interactions between these salts and zeolite are molten salt occlusion to form salt-loaded zeolite and ion exchange between the molten salt and salt-loaded zeolite. No chemical reaction has been observed. The extent of occlusion is a function of the concentration of MeCl in the zeolite and is equal to 11±1 Cl per zeolite unit cell, (AlSiO4)12, at infinite MeCl dilution. The ion-exchange mole fraction equilibrium constants (separation factors) with respect to Li are decreasing functions of concentration of MeCl in the zeolite. At infinite MeCl dilution, they are equal to 0.84, 0.87, and 2.31 for NaCl, RbCl, and CsCl, respectively, and increase in the order Na<Rb<Cs at identical MeCl concentrations. The standard ion-exchange chemical potentials are equal to −(0.0±0.5) kJ·mol−1, −(0.4±0.3) kJ·mol−1, and −(6.5±0.5) kJ·mol−1 for Na, Rb−1, and Cs+, respectively.  相似文献   

17.
It has been found that in general the volatility of dichlorides is much enhanced in the presence of gaseous A1C13 and FeCl3, and the existence of the complexes MA12C18, MAl3Cl11, and MFe2Cl8 is postulated. ΔH T, ΔS T, andT for MCl2(s) + 2AlCl3(g) = MAl2Cl8(g) are CaCl2: −17.8 kcal, −25.7 cal K−1 at 900 °K; CoCl2: −15.2, −19.4 at 750°K; MgCl2: −13.8, −17.9 at 800°K; MnCl2: −15.8, −20.9 at 750°K; NiCl2: −16.3, −24.2 at 750°K. For MCl2(s) + 3AlCl3(g) = MAl3Cl11(g) − CaCl2: −30.0, −40.5 at 900°K; CoCl2: −36.6, −47.4 at 700°K; MgCl2: −42.6, −55.4 at 750°K; MnCl2: −33.3, −42.0 at 750°K. For MCl2(s) + 2FeCl3(g) = MFe2Cl8(g) − CdCl2: −19.4, −20.9 at 700°K: CoCl2: −16.5, −17.2 at 800°K, MnCl2: −19.1, −21.2 at 750°K; NiCl2: −19.7, −24.4 at 800°K. Enhanced volatility was also found for ZnCl2, PbCl2, and CuCl, but since the condensed phase was liquid of unknown composition no calculations could be made. Owing to the interplay of the above equilibria with the dimerization equilibria for A1C13 and FeCl3 the effective vapor pressures of the dichlorides in the presence of the trichlorides pass through maxima in the region 600° to 700°C.  相似文献   

18.
High-temperature thermodynamic properties of Pt−Zr alloys containing 2 to 25 at. pct Zr and Pt−Hf alloys containing 20 to 25 at. pct Hf have been measured over the temperature range 1100 to 1400 K by a galvanic cell technique using a thoria-based electrolyte. Activities of Zr and Hf show large negative deviations from Raoult's Law; at 1300 K and 23 at. pct Zr of Hf, for instance,a Zr=6.5×10−16 anda Hf=7.9×10−17. Correlation of emf results with X-ray phase data enables calculation of standard free energies of formation of the intermetallic compounds ZrPt5, ZrPt3, and HfPt3. At 1300 K ΔG f 0 (ZrPt5) =−92,680 cal/mole; ΔG f 0 (ZrPt3)=−91,740 cal/mole; and ΔG f 0 (HfPt3)=−97,350 cal/mole. The high stabilities of phases in the Pt−Ti, Pt−Zr, and Pt−Hf systems verify the predictions of the Engel-Brewer correlation. The large negative entropies of formation of TiPt3, ZrPt3 are discussed. Applications including side reactions in fuel cells and thermocouple systems are mentioned. P. J. MESCHTER, formerly a Graduate Student at the University of Pennsylvania This paper is based upon a dissertation submitted by P. J. Meschter in partial fulfillment of the requirements of the degree of Doctor of Philosophy at the University of Pennsylvania.  相似文献   

19.
The fatigue threshold and low-rate crack propagation properties for a carbon steel, two high-strength steels, and two stainless steels were investigated in a 3 pct sodium chloride aqueous solution at frequencies between 0.03 and 30 Hz. Tests were conducted in a manner designed to avoid crack closure. Under freely corroding conditions, the effective values of the threshold stress intensity factor range, ΔKth,eff, were lower than in air for all of the steels. In particular, the ΔKth,eff values for the carbon and high-strength steels were almost equal to the theoretical ΔKth value of about 1 MPa m1/2 calculated on the basis of the dislocation emission from the crack tip. At a given ΔK level higher than the threshold, the fatigue crack propagation rates accelerated with decreasing frequency for all of the steels. Under cathodic protection, the threshold and fatigue crack propagation properties were coincident with those in air regardless of material and frequency. The observed fatigue crack propagation behavior in a 3 pct NaCl solution was closely related to the corrosion reaction of the bare surface formed at the crack tip during each loading cycle.  相似文献   

20.
Diffusion kinetics were studied by varying the time and temperature (800° to 1300°C) of dipping solid niobium into molten aluminum. It was found by X-ray diffraction analysis that only the Al3Nb phase forms at the surface of the niobium specimen and that the thickness of this layer for a given dip temperature varies parabolically with time. The activation energy and the preexponential factor of the diffusion parameter, (δD) were found to be 36.5 ∓ 0.45 kcal per mole and 2.0 ∓ 0.34 sq cm per sec, respectively.  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