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.
Vanadium nitride precipitate phase in a 9% Cr steel was observed and analyzed using transmission electron microscopy and energy dispersive spectroscopy. The steel samples were normalized at 1050 and 1100 °C for 1 h and then tempered at 750 °C for 30 min to 5 h followed by an air cooling. Through the microdiffraction pattern analyses and energy dispersive X-ray data, two kinds of vanadium nitride precipitates were determined to be (V0.6Nb0.2Cr0.2)N and (V0.45Nb0.45Cr0.1)N with the same fcc crystal structure and different lattice parameters ā = 4.070 and 4.232 Å, respectively. Lattice parameters estimated for the precipitates regarding the VN phase agree well with the present data from the microdiffraction patterns, indicating that the precipitates do not belong to the VC phase. Observed (V0.45Nb0.45Cr0.1)N precipitates consisted of undissolved particles remaining after a normalizing and the particles newly precipitated during a tempering, whilst, the observed (V0.6Nb0.2Cr0.2)N precipitates were formed during a tempering. These two vanadium nitrides seem to be a stable phase, and not an intermediary phase.  相似文献   

2.
The effect of tempering temperature on the precipitate phases in a 11%Cr ferritic/martensitic steel normalized at 1050 °C for 1 h and tempered for 2 h at temperatures ranging from 600 to 780 °C has been investigated using transmission electron microscope and energy-dispersive X-ray spectroscopy. The results show that tempering temperature does not affect the existences of niobium-rich carbonitrides, (Nb0.7V0.2Cr0.1)(C,N) and (Nb0.55V0.35Cr0.1)(C,N), vanadium-niobium-rich carbonitride (V0.45Nb0.45Cr0.1)(C,N), chromium-rich carbonitride (Cr0.83V0.12W0.05)2(C,N) and chromium-rich carbide (Cr0.7Fe0.25W0.05)23-C6, whilst the precipitations of vanadium-rich carbonitrides, (V0.65Nb0.2Cr0.15)(C,N) and (V0.55Nb0.25Cr0.2)(C,N) are dependent on tempering temperature, which were detected only at the higher tempering temperatures of 750 and 780 °C. No coarsening was occurred during the temperings for the niobium-rich and spherical vanadium-rich carbonitrides. There was a low coarsening rate for the chromium-rich carbonitrides and chromium-rich carbides with increasing the tempering temperature from 600 to 700 °C and 650 to 780 °C, respectively, and a high coarsening rate for the chromium-rich carbonitrides and chromium-rich carbides at the tempering temperatures 750 through 780 °C and 650 °C, respectively. The compositions show an increase in vanadium and a decrease in niobium and chromium contents for the niobium-rich carbonitrides, and a decrease in niobium and an increase in vanadium and chromium contents for the vanadium-niobium-rich carbonitrides, and an increase in vanadium and a decrease in tungsten contents for the chromium-rich carbonitrides. The chromium-rich carbides show an increase and a decrease in their iron and chromium contents, respectively, with increasing the tempering temperature from 650 to 780 °C.  相似文献   

3.
The precipitate phases in a 11Cr ferritic/martensitic steel normalized at 1050 °C for 1 h followed by tempered at 750 °C for 2 h have been studied through electron micro-diffraction combined with EDX analysis. Two niobium-rich carbonitride phases, (Nb0.75V0.15Cr0.1)(C,N) and (Nb0.5V0.4Cr0.1)(C,N), with the same f.c.c. structure were identified. Three vanadium-rich carbonitride phases were identified to be (V0.65Nb0.15Cr0.2)(C,N), (V0.45Nb0.45Cr0.1)(C,N) and (V0.55Nb0.25Cr0.2)(C,N) with the same f.c.c. structure. Chromium-rich carbonitride M2(C,N) and chromium-rich carbide M23C6 phases with a hexagonal and a f.c.c. crystal structure as well as a typical chemical formula of (Cr0.8V0.2)2(C,N) and (Cr0.7Fe0.25W0.05)23C6, respectively, were also identified in the steel sample. The size of each kind of precipitate phase was examined. The morphologies and precipitation sites along with the amount of observed precipitate phases were described and discussed.  相似文献   

4.
M2N nitride phases of 9% chromium steels with an extra-low carbon content have been investigated using a transmission electron microscope and an energy-dispersive X-ray (EDX) spectroscopy. The steel samples were normalized for 1 h at 1050 °C and then tempered at 600-780 °C for 30 min to 5 h followed by an air cooling. Through the analyses of the electron micro-diffraction patterns and EDX data for the precipitate particles on the extracted carbon replica, two types of Cr-rich M2N nitride phases with the same hexagonal structure but totally different lattice parameters, a = 2.80 Å/c = 4.45 Å and a = 7.76 Å/c = 4.438 Å, were determined in the steels. Four types of Cr-rich M2N phases with different lattice parameters probably existed in the steels. The M2N phase revealed a decrease in its Cr content, an increase in its V content as the tempering temperature was increased, and no obvious change in its content for the metal fraction with an increasing tempering time.  相似文献   

5.
The investigations on the precipitate phases in a 9%Cr ferritic/martensitic steel under different normalization conditions have been made by using a transmission electron microscope and an energy-dispersive X-ray spectroscopy. Hot-rolled steel samples were normalized at 1050-1200 °C for 1-2 h followed by an air cooling to room temperature. MN vanadium nitride precipitates with a plate-like morphology and a chemical formula of about (V0.4Nb0.4Cr0.2)N have been observed at triple junctions, grain boundaries and within matrix in the steel samples normalized at 1050-1150 °C for 1-2 h, but they were dissolved out at 1200 °C within 1 h. Vanadium nitride is a stable phase at 1050 °C according to thermocalc prediction of equilibrium phases in the steel. With increasing normalizing temperature and time, there was no a striking change in the chemical composition of metallic elements in the MN phase, but a considerable increase in the size of the MN precipitate.  相似文献   

6.
The on-heating phase transformation temperatures up to the melting regime and the specific heat capacity of a reduced activation ferritic-martensitic steel (RAFM) with a nominal composition (wt%): 9Cr-0.09C-0.56Mn-0.23V-1W-0.063Ta-0.02N, have been measured using high temperature differential scanning calorimetry. The α-ferrite + carbides → γ-austenite transformation start and finish temperatures, namely Ac1, and Ac3, are found to be 1104 and 1144 K, respectively for a typical normalized and tempered microstructure. It is also observed that the martensite start (MS) and finish (Mf) temperatures are sensitive to the austenitising conditions. Typical MS and Mf values for the 1273 K normalized and 1033 K tempered samples are of the order 714 and 614 K, respectively. The heat capacity CP of the RAFM steel has been measured in the temperature range 473-1273 K, for different normalized and tempered samples. In essence, it is found that the CP of the fully martensitic microstructure is found to be lower than that of its tempered counterpart, and this difference begins to increase in an appreciable manner from about 800 K. The heat capacity of the normalized microstructure is found to vary from 480 to 500 J kg−1 K−1 at 500 K, where as that of the tempered steel is found to be higher by about, 150 J kg−1 K−1.  相似文献   

7.
Phase equilibria in the system Si-U-V were established at 1100 °C by optical microscopy, EMPA and X-ray diffraction. Two ternary compounds were observed, U2V3Si4 and (U1−xVx)5Si3, for which the crystal structures were elucidated by X-ray powder data refinement and found to be isotypic with the monoclinic U2Mo3Si4-type (space group P21/c; a = 0.6821(3), b = 0.6820(4), c = 0.6735(3) nm, β = 109.77(1)°) and the tetragonal W5Si3-type (space group I4/mcm, a = 1.06825(2), c = 0.52764(2) nm), respectively. (U1−xVx)5Si3 appears at 1100 °C without any significant homogeneity region at x ∼ 0.2 resulting in a formula U4VSi3 which corresponds to a fully ordered atom arrangement. DTA experiments clearly show decomposition of this phase above 1206 °C revealing a two-phase region U3Si2 + V3Si. At 1100 °C U4VSi3 is in equilibrium with V3Si, V5Si3, U3Si2 and U(V). At 800 °C U4VSi3 forms one vertex of the tie-triangle to U3Si and V3Si. Due to the rather high thermodynamic stability of V3Si and the corresponding tie-lines V3Si + liquid at 1100 °C and V3Si + U(V) below 925 °C, no compatibility exists between U3Si or U3Si2 and vanadium metal.  相似文献   

8.
We report on the crystallographic aspects and the basic properties of the plutonium based compound PuPd5Al2. This material is antiferromagnetic at TN = 5.6 K and does not present any hint of superconductivity down to 2 K. This material crystallizes in the ZrNi2Al5-type of structure with lattice parameters: a = 4.1302 Å and c = 14.8428 Å. The magnetization, heat capacity and electrical resistivity measurements indicate clearly antiferromagnetic order at TN = 5.6 K. This material is compared to the structurally related cerium based material CePd5Al2 presenting superconductivity induced by pressure.  相似文献   

9.
The microstructure of thermally grown oxides (TGO) and the creep properties of alloy 617 were investigated. Oxidation and creep tests were performed on 100 μm thick foils at 800-1000 °C in air environment, while the thickness of TGO was monitored in situ. According to energy dispersive X-ray (EDX) mapping micrographs observation, superficial dense oxides, chromia (Cr2O3), which was thermodynamically unstable at 1000 °C, and discrete internal oxides, alumina (α-Al2O3), were found. Consequently, the weight of the foil specimen decreased due to the spalling and volatilization of the Cr2O3 oxide layer after an initial weight-gaining. Secondary and tertiary creeps were observed at 800 °C, while the primary, secondary and tertiary creeps were observed at 1000 °C. Dynamic recrystallization occurred at 800 °C and 900 °C, while partial dynamic recrystallization at 1000 °C. The apparent activation energy, Qapp, for the creep deformation was 271 kJ/mol, which was independent of the applied stress.  相似文献   

10.
Rapid synthesis of Pb5(VO4)3I, a potential immobilisation host for iodine radioisotopes, was achieved in an open container by microwave dielectric heating of a mixture of PbO, PbI2, and V2O5 at a power of 800 W for 180 s (at 2.45 GHz). The resulting ceramic bodies exhibited a zoned microstructure, differentiated by inter-granular porosity and phase assemblage, as a consequence of the inverse temperature gradient characteristic of microwave dielectric heating. Liquid PbI2 within the interior of microwave processed ceramics assisted formation of Pb5(VO4)3I, and reduced inter-granular porosity. In contrast, the exterior of microwave processed ceramics comprised poorly sintered Pb5(VO4)3I with the presence of minor reagent relics. Quantitative microanalysis, electron diffraction and Rietveld analysis, confirmed the synthesis of stoichiometric Pb5(VO4)3I within precision. The crystal structure of Pb5(VO4)3I was found to adopt space group P63/m with a = 10.4429(3) Å and c = 7.4865(2) Å.  相似文献   

11.
Weld beads on plate specimens made of type 316L stainless steel were neutron-irradiated up to about 2.5 × 1025 n/m2 (E > 1 MeV) at 561 K in the Japan Material Testing Reactor (JMTR). Residual stresses of the specimens were measured by the neutron diffraction method, and the radiation-induced stress relaxation was evaluated. The values of σx residual stress (transverse to the weld bead) and σy residual stress (longitudinal to the weld bead) decreased with increasing neutron dose. The tendency of the stress relaxation was almost the same as previously published data, which were obtained for type 304 stainless steel. From this result, it was considered that there was no steel type dependence on radiation-induced stress relaxation. The neutron irradiation dose dependence of the stress relaxation was examined using an equation derived from the irradiation creep equation. The coefficient of the stress relaxation equation was obtained, and the value was 1.4 (×10−6/MPa/dpa). This value was smaller than that of nickel alloy.  相似文献   

12.
Ten types of 23Na implanted targets have been fabricated for the purposes of investigating the effects of proton beam bombardment on the implanted sodium distribution. Targets were implanted at energies of ENa = 10-30 keV using copper, tantalum, and nickel as host materials. Thin layers (100-200 Å) of chromium and gold were also evaporated over some of the targets to provide a protective layer for the implanted sodium. The 23Na(pγ) resonance at a lab proton energy of Ep = 309 keV was used to determine the implanted distribution. Successive resonance profile measurements are presented for each implanted target, and the concurrent loss of 23Na resulting from beam bombardment is reported. The calculated temperature rise of the targets indicates that beam heating has a negligible effect on the implanted sodium distribution, and that the principal mechanism for 23Na loss during beam bombardment is sputtering.  相似文献   

13.
The total mass attenuation coefficients (μm), for Cr, Fe, Ni and FexNi1−x (x = 0.8, 0.7, 0.6, 0.5, 0.4, 0.3 and 0.2), FexCryNi1−(x+y) (x = 0.7, y = 0.1; x = 0.5, y = 0.2; x = 0.4, y = 0.3; x = 0.3, y = 0.3; x = 0.2, y = 0.2 and x = 0.1, y = 0.2) and NixCr1−x (x = 0.8, 0.6, 0.5, 0.4 and 0.2) alloys were measured at 22.1, 25.0, 59.5 and 88.0 keV photon energies. The samples were irradiated with 10 mCi Cd-109 and 100 mCi Am-241 radioactive point source using transmission arrangement. The γ- and X-rays were counted by a Si(Li) detector with a resolution of 160 eV at 5.9 keV. Total atomic and electronic cross-sections (σt and σe), effective atomic and electron numbers (Zeff and Neff) were determined experimentally and theoretically using the obtained mass attenuation coefficients for investigated 3d alloys. The theoretical mass attenuation coefficients of each alloy were estimated using mixture rule. The experimental values were compared with the calculated values for all samples.  相似文献   

14.
The goal is to evaluate the neutronic behavior when (Pu–U) and (Am–Pu–U) mixed oxide are inserted in a typical cell of a Pressurized Water Reactor (PWR) such as Angra-I. Four types of fuels were studied: (1) MOX fuel enriched at 3.1% and Vm/Vf = 1.15; (2) MOX fuel enriched at 4.5% and Vm/Vf = 1.15; (3) MOX fuel enriched at 4.5% and Vm/Vf = 2.0 and (4) MOX fuel enriched at 4.5%, with 1% of Americium insertion in its composition (62.8% of Am241, 0.1% of Am242m and 37.1% of Am243) and with Vm/Vf = 2.0. The first case represents the standard state of Angra I, but with Pu. The second case is similar to the first but the enrichment is increased. To evaluate the Americium insertion, a study of the Vm/Vf was made and better results were obtained with Vm/Vf = 2.0 and to compare, this case was too evaluated to (Pu–U) in the third and fourth cases. The idea is to verify the possibility of using these fuels in Angra-I analyzing neutronic parameters such as infinite multiplication factor, hardening spectrum, Boron worth and reactivity temperature coefficients. The results show that it is possible to use all the studied fuels in Angra-I as well as to burn Am inserted in the MOX fuel by a considerable quantity during PWR operation. The WIMS-D5 code was used to perform a simplified neutronic and burnup simulations to evaluate this possibility.  相似文献   

15.
The fracture properties of the tempered martensitic steel Eurofer97, which is among the main candidates for fusion power plant structural applications, were studied with two sizes of pre-cracked compact specimens (0.35T C(T) and 0.87T C(T)). The fracture toughness behavior was characterized within the temperature range −80 to −40 °C. The ductile-to-brittle transition reference temperature, as defined in the ASTM standard E1921, was around T0 ≈ −75 °C. At −60 °C, it was found that two sets of toughness data obtained with 0.35T and 0.87T C(T) specimens are not consistent with the size adjustments recommended in the ASTM standard. It was then shown that the underlying reason of this inconsistency is an inappropriate specimen size limit of the ASTM standard for this type of steel. From published fracture toughness data on the tempered martensitic steel F82H steel, similar results were also highlighted. 3D finite elements simulations of the compact specimens were performed to compare the stresses and deformations at the onset of fracture. A local approach model based on the attainment of a critical stress and a critical volume was used to study the constraint loss phenomenon. Within the framework of this model, the strong toughness increase by reducing the specimen size could be satisfactorily explained.  相似文献   

16.
The oxidation of iron and chromium that are present as impurities in zirconium metal or as alloying elements in Zircaloy-4 was investigated with PhotoElectroChemical techniques (PEC), highlighting the chemical nature, the size and the lateral distribution of Fe and Cr-containing phases in thin zirconia scales formed during the oxidation of pure zirconium and Zircaloy-4 at 470 °C in oxygen. In the case of zirconium, iron and chromium impurities led to the formation of oxides distributed in a homogeneous way in the zirconia scale, while in the case of Zircaloy-4 these elements, present in the form of intermetallic particles in the substrate, led to the formation of localised haematite Fe2O3, rhomboedric solid solution (FexCr1−x)2O3 and chromia Cr2O3 phases. These phases were accurately studied via the measurement of their respective band-gap (Fe2O3: 2.2 eV, (FexCr1−x)2O3: 2.6 eV and Cr2O3: 3.0 eV). It is concluded that PEC techniques represent a sensitive and powerful way to locally analyse the various semiconductor phases in the oxide scale at a micron scale.  相似文献   

17.
The new ternary compound UFe5Ga7 was prepared by an argon arc-melting followed by annealing. This intermetallic compound belongs to the series UFexGa12−x and crystallizes in a structure related with the ThMn12-type, with a = 8.6309 Å and c = 5.0524 Å. EDS elemental analysis yielded the phase composition UFe4.9(2)Ga6.8(1), which is very close to the nominal composition. Dc magnetization measurements revealed ferromagnetic type of magnetic ordering in UFe5Ga7. The Curie temperature of ∼439 K was estimated by the temperature dependence of magnetization at a low magnetic field of 0.1 T.  相似文献   

18.
A new phase PuPdSn was prepared and studied by X-ray diffraction, magnetization and heat capacity measurements, performed in the temperature range 2-300 K and in magnetic fields up to 14 T. The crystal structure determined from single-crystal X-ray data is the hexagonal ZrNiAl-type [space group ] with lattice parameters: a = 7.5057 Å and c = 4.0853 Å. PuPdSn orders antiferromagnetically at TN = 21 K. Moreover, another antiferromagnetic-like transition takes place at 9.6 K. Above TN the susceptibility follows a modified Curie-Weiss law with μeff = 1.0 μB, Θp = −14 K and χ0 = 2.1 × 10−4 emu/mol. The low-temperature linear specific heat coefficient is small (γ ∼ 8 mJ/mol K2) pointing to well localized 5f electrons.  相似文献   

19.
During service, Zr-2.5Nb pressure tubes of nuclear power reactors may be prone to suffer from crack growth by delayed hydride cracking (DHC). For a given hydrogen plus deuterium concentration there is a critical temperature (TC) below which DHC may occur. In this work, TC was measured for specimens cut from pressure tubes made in Canada (CANDU) and in Russia (RBMK). Hydrogen was added to the specimens to get concentrations ranging from 24 to 60 wt ppm. It was found that TC was higher than the corresponding precipitation temperature. The crack propagation velocity (VP), measured in axial direction, increases from a minimum at TC to a maximum at a temperature close but higher than the precipitation temperature. At lower temperatures, when hydride precipitates are present in the bulk, VP follows an Arrhenius law: VP = A exp(−Q/RT), with an activation energy Q of 66-68 kJ/mol for both tubes. The RBMK material presented lower velocities than CANDU one.  相似文献   

20.
Studies were conducted on the creep behavior of Alloy 800H in impure helium and in a 1%CO-CO2 environment. At relatively low applied stresses and at low temperatures, the presence of methane in helium reduced the rupture strain significantly while increasing the rupture life relative to the behavior in pure helium. The degradation in rupture strain is due to the occurrence of cleavage fracture in the He + CH4 environment; this explanation is also supported by high activation energy (Q = 723 kJ/mol) for creep in He + CH4. At higher applied stresses and also at higher temperatures, creep-rupture behavior in He and He + CH4 was similar. Creep response in pure He and in CO-CO2 follows a dislocation climb-controlled power-law behavior whereas that in He + CH4 has a different behavior as indicated by the high stress exponent (n = 9.8). The activation energy for creep in pure He was 391 kJ/mol and in CO-CO2 was 398 kJ/mol, and appeared to be independent of stress in both environments. On the other hand, in He + CH4, the activation energy (Q = 723 kJ/mol) seems to be dependent on stress.  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