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.
The structural relationship between the hydride phases in Ti–Mo–H solid solution system (Mo content up to 15 at% in the alloy) during dehydrogenation process under annealing has been studied by conventional and in situ X-ray powder diffraction and transmission electron microscopy (TEM) analysis. During dehydrogenation, the saturated hydrides of the Ti–Mo alloys with fcc δ-phase structure transfer into bcc β-phase at higher temperatures. An associated hydrogen concentration reduction for the δ-phase hydride is observed in the process. However, as the hydrogen concentrations decrease to certain values (H/M  1.1–1.7), the unsaturated δ-phase formed at high temperature would become unstable at lower temperature, and transfer into a tetragonal phase (denoted the -phase here). Unlike that of the -phase in Ti–H system, the phase transition does not occur for the saturated δ-phase with hydrogen concentration close to the stoichiometric limit. The hydrogen concentration of this -phase hydride is in between that of the tetragonal γ and -phase in Ti–H system, but more close to the γ-phase. The occurrence region of this -phase expands along with the increase of the Mo content in the alloys. The phase has a lattice similar to that of the -phase in Ti–H system with corresponding fct unit-cell c/a < 1.  相似文献   

2.
The well-established “halide salt” route was employed in the present work to produce Al–Ti–B grain refiner alloys with consistent, good properties. The holding step in the production cycle was revised, however, to avoid oxidation of the molten alloy which is believed to be responsible for the relatively low Ti recoveries and thus for the inadequate and inconsistent grain refining efficiency. Stirring during holding was found to degrade the grain refining properties when molten potassium aluminium fluride salt was left on the molten alloy to avoid excessive oxidation. Likewise, holding temperatures higher than 800 °C and holding times longer than 30 min both had an undesirable effect on the grain refining performance. The experimental Al–5Ti–1B grain refiner alloy produced according to the present method provided consistent and better overall grain refining performance.  相似文献   

3.
Amorphous alloys are anticipated as new membrane materials for high purity hydrogen production, as substitutes for expensive palladium alloys. For amorphous Zr–Ni-based alloys reported to date, hydrogen permeability increases with Zr content. Hydrogen solution properties in a series of amorphous Zr–Hf–Ni ternary alloys were measured carefully using the Sieverts method and residual hydrogen measurements to investigate the reason. Results indicate that hydrogen solubility in the ternary alloys increases with increasing Zr to improve hydrogen permeability, not because of the geometrical atomic structure but because of higher hydrogen affinity of Zr than that of Hf. Increased permeability with Zr in other amorphous Zr–Ni-based alloys is also expected to be attributable to the same reason. Additionally, hydrogen was found with low mobility, and was not removable even after 10 h evacuation at 573 K; the importance of decreasing low mobility hydrogen as a countermeasure against hydrogen embrittlement was pointed out. Equilibrium hydrogen concentration was found not to obey Sieverts’ law with respect to hydrogen pressure. Rather, it was linear roughly to the quarter power. Parameters to reproduce pressure–composition isotherms were determined using Kirchheim's theory.  相似文献   

4.
The effect of trace levels of Ni on the intermediate temperature creep behavior of the alloy Ti–6Al–2Sn–4Zr–2Mo (wt%) has been investigated. Creep experiments were performed in tension over the temperature range 510–565 °C at stress range 138–413 MPa. Two heats of commercial grade Ti–6Al–2Sn–4Zr–2Mo with Ni levels of 0.006 and 0.035 wt% were studied. The high Ni material uniformly exhibited higher primary creep strains and minimum strain rates than the lower Ni material. Stress exponents in the range 5–7 and 4–6 were obtained for the high Ni and low Ni material respectively. At 565 °C a transition to a low stress region with a stress exponent equal 1 is found for both materials. At all stress levels, the apparent activation energy was lower for the high Ni material. The apparent activation energy is in excellent agreement with those reported for lattice self-diffusion in -titanium in the presence of fast diffusing impurities. The results also suggest that creep in the higher stress regime is controlled by dislocation motion within the -phase. We suggest that trace levels of Ni in the -phase accelerate self-diffusion therefore increasing the rate of dislocation climb leading to the higher creep rates observed in the high Ni material. In Part II, direct evidence in support of dislocation-based creep being important in both low and high stress regimes is presented.  相似文献   

5.
6.
Laser gas assisted processing can be used to modify the surface properties of Ti–6Al–4V alloy through the use of gaseous interaction with the laser melted surface. Laser surface melting of titanium and its alloys in nitrogen to form a layer of TiN embedded in a metallic matrix which is enriched in alloying elements has attracted considerable interest. The surface roughness of the laser-treated surface is poor, therefore, a secondary processing becomes essential. In the present study, duplex treatment of Ti–6Al–4V alloy was carried out. The alloy surface was melted initially under a controlled nitrogen atmosphere, which in turn resulted in a laser-induced nitrided surface. The resulting workpiece surface, then, was PVD TiN coated. In order to assess the wear properties of the resulting surface, friction tests were carried out. SEM, XRD and microhardness were carried out for microstructural analysis and material characterization. It was found that the adhesion of the TiN coating to the base alloy improved considerably in the case of laser-treated workpieces and smooth transition in plastic shearing resistance between the TiN coating and the base alloy enhanced the wear properties of the laser-treated surface.  相似文献   

7.
The hydrogen absorption properties of Ti–0.2 mass% Pd (Ti–0.2Pd) alloy in 2.0% and 0.2% acidulated phosphate fluoride (APF) and neutral 2.0% NaF solutions (25 °C) has been evaluated by hydrogen thermal desorption analysis. During the early stage of immersion (120 h) in the 2.0% APF solution, the amount of absorbed hydrogen was lower than 500 mass ppm. A thermal desorption of hydrogen primary appearing with a peak at 500–600 °C and a broad desorption ranging from 100 to 400 °C were observed. In the 0.2% APF solution, the amount of absorbed hydrogen saturated at 100–200 mass ppm; the thermal desorption of hydrogen appeared with a single peak at 550 °C. In the 2.0% NaF solution, hydrogen absorption was negligible even after 1000 h of immersion, although corrosion pits were observed. The results of the present study suggest that the hydrogen absorption of Ti–0.2Pd alloy, as compared with commercial pure titanium, is suppressed in fluoride solutions.  相似文献   

8.
The effect of heat treatment on microstructure, equilibrium hydrogen sorption pressure and plateau slope of LaNi3.8Al0.75Mn0.45 alloy was investigated. X-ray diffraction (XRD) analysis indicates that annealed alloys have single phase and the same hexagonal structure as that of LaNi5 alloy (CaCu5 type, P6/mmm ). The cell parameters of alloys fluctuate with the increasing annealing temperature. The equilibrium hydrogen pressure and plateau slope are a parabola function with annealing temperature for LaNi3.8Al0.75Mn0.45 alloy. By this relationship, an appropriate heat treatment temperature for LaNi3.8Al0.75Mn0.45 alloy is determined to about 1220-1230 K by mathematic simulation process. However, the maximum hydrogen storage capacity of alloys does not affected by the annealing temperature.  相似文献   

9.
The variability in fatigue life of the Ti–6Al–2Sn–4Zr–6Mo (Ti-6-2-4-6) alloy was investigated. Cumulative life distribution plots were found to be composed of two failure mechanisms. The data could be closely represented by a cumulative distribution function (CDF) resulting from the superposition of the CDFs of the individual mechanisms. An approach for life prediction based on the data due to the worst-case mechanism is suggested.  相似文献   

10.
Using a composite oscillator, the Young’s modulus and the internal friction of annealed, plastically bent and bent–straightened V–4Ti–4Cr alloy specimens have been measured, in a wide amplitude range, before, during and after proton irradiation. The proton energy and flux were 8 MeV and 1012 p cm−2 s−1. The in situ dependencies on the proton beam current and temperature have been obtained. Softening and overheating (which looks like softening–strenghtening of the alloy) effects, which arise due to radiation damage, have been revealed. The influence of beam heating and defect accumulation–annihilation on the acoustic properties of the annealed and pre-strained samples, during and after proton irradiation, are briefly discussed.  相似文献   

11.
It is of great interest to replace the K2TiF6 salt so as to reduce the volume of fluoride-bearing particulate material to be added to the aluminium melt in the popular “halide salt” process. Ti sponge was used in the present work as the source of Ti in the production of an Al–5Ti–1B grain refiner. Addition of Ti granules into molten aluminium, either premixed with or before KBF4 salt, has produced Al–5Ti–1B alloys where the boride particles were relatively few and predominantly of the AlB2 type. The grain refining efficiency of these alloys were far from satisfactory. TiB2 was the dominant boride phase with sufficient number of blocky aluminide particles when Ti, in excess of the TiB2 stoichiometry was supplied before hand and the balance was reserved for co-addition with KBF4. Al3Ti particles were generated soon after the Ti solubility limit was exceeded in the first step while the boride particles were subsequently produced by the reaction between molten aluminium, KBF4 and K2TiF6. The Al–5Ti–1B master alloy thus produced provided an adequate grain refining performance while the amount of particulate material to be added to the aluminium melt was reduced by nearly 30%.  相似文献   

12.
The master alloys based on the Al–Ti–B system have been used extensively for refining the grain structure of aluminum alloys. The quality-related problems linked with the TiB2 particles, however, have generated an interest in the Al–Ti–C grain refiners as an acceptable replacement for Al–Ti–B master alloys. TiC particles are smaller than the TiB2 particles and are less prone to agglomeration. Al–3Ti–0.15C grain refiners have been in use for some time in several alloy systems. Much of the work reported on this alloy, however, has been from DC casting while performance data in strip casting is not available. In the present work, a commercial Al–3Ti–0.15C grain refiner was employed in the twin-roll casting of AA8111 foilstock. Its grain refining efficiency was compared with that of the Al–5Ti–0.2B master alloy, the standard grain refiner in aluminium industry for the manufacture of aluminium foil products.  相似文献   

13.
There are strong needs for productive/quality machining strategies of notoriously “difficult-to-machine” aerospace materials. The current means of machining these materials is dominated by mechanical cutting methods, which are costly due to high tooling costs, poor surface quality and limitations in the workpiece features and operations that can be machined. The newest EDM technology may be able to circumvent problems encountered in mechanical machining methods. In this paper, the EDM technology has been used to machine titanium alloy Ti–6Al–4V to investigate the effect of Ti–6Al–4V's thermal and electrical properties on the EDM productivity. In the study, temperature measurements have been made for Ti–6Al–4V workpieces with various duty factors to clarify the essential causes of difficulty in machining titanium alloys and observe the optimal duty factor in terms of productivity and quality.  相似文献   

14.
The mechanical alloying technique was used to make multi-phase alloys V TiFe0.85Mn0.15.Their hydrogen storage properties were characterized and compared with that of the polycrystalline alloys made by casting.It was found that the ball milled alloys can absorb hydrogen at room temperature in the first cycle without prior activation.The 40% V 60% TiFe0.85Mn0.15 alloy made by mechanical alloying shows the best hydrogen storage property with the valid hydrogen capacity up 50 220mL/g at 293K and 4.0MPa, and the P-C-T behavior is also improved.The XRD and EDX analyses also show that the phase of these alloys contains.V,TiFe,γ-TiMnx,TiFe2 and α-FeV.The composition of these phases affects significantly the hydrogen storage properties of alloys.  相似文献   

15.
The impact of the salt addition practice on the microstructure and grain refining efficiency of Al–Ti–B alloys produced by the “halide salt” route was investigated. The grain refining performance of an experimental Al–5Ti–1B master alloy was optimized when the halide salts were pre-mixed before addition to aluminium melt at 800 °C during the production of the grain refiner. The stirring action provided during salt addition was found to degrade, while a high rate of addition was found to improve, the grain refining efficiency. In view of the above, an improved salt addition practice to ensure an exceptional grain refining performance is claimed to comprise the following steps: melting commercial purity aluminium ingot; addition of pre-mixed salts to molten aluminium at 800 °C, at once to facilitate a rapid salt reaction, gently mixing the salts with the aluminium melt without introducing any stirring. The grain refiner master alloy thus produced gives an average grain size of 102 μm 2 min after inoculation.  相似文献   

16.
Yin-Chih Lin 《Acta Materialia》1999,47(18):4665-4681
Microstructures and superparamagnetic properties in aged-hardened Fe–9%Al–30%Mn– (x)C,Si alloys, resulting from overaging at a temperature of 823 K for 48 h to 313 days, have been investigated by transmission electron microscopy (TEM), X-ray diffraction patterns, and vibrating sample magnetometry (VSM). The results reveal that the precipitate κ-phase [(Fe,Mn)3AlC] decomposition in this alloy, overaged at 823 K for one week, resulted from two separate mechanisms: (1) wetting of the antiphase boundary segment (APBs) of D03 [(Fe/Mn)3Al] domains by the B2 [(Fe/Mn)Al] phase; and (2) precipitation of the B2 [(Fe/Mn)Al] phase within the domain. A superparamagnetic behaviour was discovered when the alloy was overaged at 823 K for ≈120–313 days. The super-soft magnetic property was mainly attributable to the ferromagnetic spinel-ordered (B2 [(Fe/Mn)Al]+D03 [(Fe/Mn)3Al]) phases and ordered B2 with monoclinic ′Mn structures.  相似文献   

17.
A Ti3Al-based titanium aluminide alloy, Ti–24Al–11Nb, was cathodically charged with hydrogen in a 5% H2SO4 aqueous solution for various charging times, and the formation and dissociation of the hydride, the hydrogen evolution behavior and the total hydrogen uptake were investigated mainly by means of X-ray diffractometry and thermal desorption spectroscopy (TDS). The same kind of hydride phase as observed previously in Ti–25Al alloy (hexagonal hydride) was presumably formed in the Ti–24Al–11Nb alloy after cathodic charging. No damage, such as cracks, was induced by hydrogen charging. Two kinds of TDS peaks, one probably corresponding to hydride dissociation and the other to hydrogen dissolution in the normal lattice site, were found after longer hydrogen charging. It is suggested that niobium addition to Ti3Al-based titanium aluminide alloy may reduce hydrogen susceptibility during cathodic charging.  相似文献   

18.
The hydrogen absorption–desorption performance of the body-centered-cubic (bcc) Ti–V–Cr–Pd alloys have been investigated. Ti33V33Cr34 ingots with 0, 0.05, 0.5 at.% Pd were prepared by arc melting. X-ray diffraction (XRD) revealed that all of these alloys were homogeneous bcc solid solutions. Pd-containing (0.05, 0.5 at.% Pd) Ti–V–Cr alloys have better initial activation properties than those without Pd, and the desorption plateau pressure of the (Ti33V33Cr34)99.5Pd0.5 alloy was substantially higher than that of the alloy without Pd. It is also found that the hysteresis difference is smaller in these alloys and degradation of hydrogen absorption capacity becomes steady after the 25th cycling test. (Ti33V33Cr34)99.5Pd0.5 alloy exhibits large hydrogen absorption and desorption capacity of up to 3.42 and 2.07 mass% at 353 K, respectively.  相似文献   

19.
It is very desirable to replace the KBF4 salt in the popular “halide salt” process to reduce the volume of fluoride salts to be added to molten aluminium in the production of Al–Ti–B grain refiners. Being over 2 times richer in B, Na2B4O7 is a promising replacement for KBF4, and is used in the present work to produce Al–Ti–B grain refiner master alloys. A fraction of the aluminide particles were entrapped in the spent salt giving a relatively lower Ti recovery when KBF4 was replaced by Na2B4O7. The grain refining performance of the Al–Ti–B grain refiner alloy thus produced was nevertheless acceptable. The spent salt became too viscous with the oxides, aluminides and borides to be removed by decanting when Na2B4O7·5H2O was used to supply boron. The viscous spent salt, entrained in the grain refiner alloy, did not only impair its performance, but also hurt the fluidity of the molten alloy and made pouring difficult.  相似文献   

20.
The effect of manganese addition on hydrogen storage performance of vanadium-based BCC alloys was investigated by measuring mainly pressure-composition (P−C) isotherms at 303K. Annealing heat-treatment was also considered in selected cases. The XRD patterns showed BCC single phase in all the alloys. With increasing Mn content, the lattice parameters decreased linearly, thus resulting in an increase in plateau pressure and a reverse effect on maximum hydrogen storage capacity. However, an effective hydrogen storage capacity as high as 1.92 wt.% was achieved at x=0.075. V−Ti−Mn alloys showed a surprisingly flat desorption plateau, but lots of absorbed hydrogen cannot desorb at ambient temperature. Although V0.44Ti0.20Cr0.12Mn0.12Fe0.12 alloy did not show the first plateau, both the maximum and effective hydrogen storage capacities were very low.  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