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.
Zr-based bulk metallic glasses (BMGs) exhibit interesting mechanical properties since they combine high fracture stress, elastic strain (up to 2%), significant fracture toughness and good corrosion resistance. Quaternary systems with general composition Zr–Ni–Cu–Ti show wide composition ranges in which BMG can be obtained. The addition of the another element to the quaternary alloys often increases the glass forming ability (GFA). The aim of this work was to study the influence of aluminium content on the GFA and on the mechanical properties of the Zr–Ni–Cu–Ti alloys. Multicomponent Zr75−xAlxNi10Cu10Ti5 (x = 15, 20 at%) alloys were produced by melt spinning method obtaining ribbons, and by casting technique into a copper mould, manufacturing rod shape samples with maximum diameter of 2 mm. Supercooled liquid region depends on chemical composition and exceeds 45 °C. Vickers microhardness of studied alloys is comparable to the highest ones for other Zr-based BMG.  相似文献   

2.
Nuclear magnetic resonance (NMR) has been used for the first time to directly monitor the dynamic partitioning of Cu atoms from shearable precipitates into the solid solution as a function of straining at room temperature in two Al–Cu-based alloys. Al–3Cu–0.05Sn (wt.%) and Al–2.5Mg–1.5Cu (wt.%) alloys were heat-treated to provide a fine distribution of 5 nm Guinier–Preston (GP) zones and <1 nm Guinier–Preston–Bagaryatsky (GPB) zones, respectively, and were then subjected to rolling strains up to 100%. It is shown that in the Al–Cu–0.05Sn alloy, strains up to 40% can pump solute from the 5 nm GP zones back into solid solution for the temperature and strain-rate of deformation employed here. In the case of the Al–Cu–Mg alloy, no dissolution of the GPB zones is observed. A simple model for the strain-induced dissolution of the shearable precipitates is given and compared with the experimental results. The dependence of the Cu repartitioning process on the precipitate size is emphasized. These observations and modeling give guidelines for the design of Al–Cu-based alloys to exploit the dynamic interplay of strain-induced Cu partitioning between metastable states, e.g. solid solution and GP (or GPB) zones, for tailoring ultimate mechanical properties. It is proposed that this strain-induced phase transformation is a form of dynamically responding microstructure that can be employed to obtain aluminum alloys with well-designed microstructures.  相似文献   

3.
Binary Mg–Cu amorphous alloys were first fabricated in 1980s via liquid quenching. In this study, the Mg1−xCux (x varying from 38 at.% to 82 at.%) partially amorphous thin films are prepared via co-sputtering. Upon thermal annealing, the Mg2Cu or MgCu2 nanocrystalline phases are induced in the Mg-rich or Cu-rich thin films, respectively. Due to the presence of fine nanocrystalline Mg2Cu or MgCu2 particles in the Mg–Cu amorphous matrix, the as-sputtered thin films show satisfactory Young's modulus 100 GPa and hardness 4 GPa.  相似文献   

4.
In order to improve the thermoelectric properties via efficient phonon scattering Didymium (DD), a mixture of Pr and Nd, was used as a new filler in ternary skutterudites (Fe1−xCox)4Sb12 and (Fe1−xNix)4Sb12. DD-filling levels have been determined from combined data of X-ray powder diffraction and electron microprobe analyses (EMPA). Thermoelectric properties have been characterized by measurements of electrical resistivity, thermopower and thermal conductivity in the temperature range from 4.3 to 800 K. The effect of nanostructuring in DD0.4Fe2Co2Sb12 was elucidated from a comparison of both micro-powder (ground in a WC-mortar, 10 μm) and nano-powder (ball-milled, 150 nm), both hot pressed under identical conditions. The figure of merit ZT depends on the Fe/Co and Ni/Co-contents, respectively, reaching ZT > 1. At low temperatures the nanostructured material exhibits a higher thermoelectric figure of merit. The Vickers hardness was measured for all samples being higher for the nanostructured material.  相似文献   

5.
Coincidence doppler broadening (CDB) spectroscopy has been applied to study the precipitation process induced by aging in Mg–1.0 wt.% Ca and Mg–1.0 wt.% Ca–1.0 wt.% Zn alloys. In addition positron lifetime experiments and microhardness measurements have been performed. A peak centered at 11.5 × 10−3m0c is found in the CDB ratio spectra of the alloys aged at 473 K. It is attributed to annihilations with the core electrons of Ca. The results indicate the formation of a particle dispersion that hardens the alloys. This dispersion is correlated with the appearance of the peak attributed to Ca atoms. Zn atoms in the Mg matrix inhibit the formation of quenched-in vacancies bound to Ca atoms in the aged ternary alloy producing the dispersion refinement.  相似文献   

6.
M.C. Lin  C.Y. Tsai  J.Y. Uan   《Corrosion Science》2009,51(10):2463-2472
This study investigated the electrochemical and corrosion performance of Mg–Li–Al–Zn anodes with Al compositions of 3 wt.% and 9 wt.%. Mg–Li–Al–Zn alloy with 9 wt.% Al had a relatively negative open-circuit potential and a high discharge voltage in MgCl2 electrolyte, owing to the distribution of numerous AlLi particles in the matrix of the alloy. AlLi particles were believed to transform to Al particles during the corrosion of the Mg–Li–Al–Zn anode. The high-Al anode material exhibited good corrosion performance since a dense and continuous Mg(OH)2/Al composite layer covered the surface of the high-Al anode. Experimentally, increasing the Li+ concentration in the electrolyte improved the corrosion performance of the Mg anode.  相似文献   

7.
Mg65Cu25Re10 (Re = Y, Gd) and Mg64Cu25Nb1Re10 (Re = Y, Gd) bulk metallic glasses (BMGs) have been fabricated by copper mould casting. It is shown that the minor addition of Nb can not only increase the thermal stability but also improve the fracture strength and the toughness of the Mg-based BMG alloys greatly. The fracture strength of Mg64Cu25Nb1Y10 and Mg64Cu25Nb1Gd10 reaches as high as 1023 MPa and 973 MPa, respectively. The Young's modulus has also been increased by the addition of Nb. From the fracture morphologies of the Nb bearing Mg-based BMG alloys, it is known that the size of plastic deformation zone can be increased to micrometer scale. This work proves that it is possible to improve the strength and toughness of Mg-based BMG by adding an element having positive heat of mixing with the constituent elements.  相似文献   

8.
Three isopleths at the Mg-rich corner of Mg–Mn–Ce ternary system were investigated via thermal analysis, SEM/EPMA and XRD. A ternary eutectic reaction was observed at 1 wt.% Mn and 23 wt.% Ce and 592 °C. A solid-solution type ternary intermetallic compound, (Mg,Mn)12Ce, was observed with 0.5 at% solid solubility of Mn in the tetragonal Mg12Ce. With the aid of thermodynamic modeling and experiments, a revised phase diagram for the binary Mg–Ce system and the isopleths of 0.6, 1.8 and 2.5 wt.% Mn were proposed up to 25 wt.% Ce.  相似文献   

9.
Composites with ferromagnetic nanoparticles, Fe and Fe50Ni50, dispersed in Al2O3 have been synthesized by a solution phase technique. The structure and magnetic properties of these composites with varying fractions of Al2O3 have been investigated. Both Fe and Fe50Ni50 nanoparticles are amorphous in the as-prepared state and become crystalline on heat treating with near equilibrium lattice parameters of 0.287 nm and 0.358 nm respectively. The interparticle distance increases with increasing Al2O3 from 0 wt.% to 20 wt.%. The size of Fe nanoparticles is 40 nm while the Fe50Ni50 nanoparticles are 20 nm in size. The Fe and Fe50Ni50 nanoparticles dispersed composites are found to be ferromagnetic at room temperature both in the as-prepared and heat treated conditions with clear coercive fields of 5.5–35 × 103 A m−1. The saturation magnetization increases by orders of magnitude on heat treatment, for e.g. from <1.0 emu g−1 to 143.4 emu g−1 for Fe–15 wt.% Al2O3 and 95.6 emu g−1 for Fe50Ni50–15 wt.% Al2O3. The Fe-composites exhibit a Curie transition at 1000 K while the Fe50Ni50 composites exhibit a transition at 880 K, both temperatures close to bulk values.  相似文献   

10.
Structural, thermodynamical and magnetic properties of Fe73.5−xSi13.5B9Cu1Nb3Mnx amorphous alloys, with Mn content x=1,315, were studied by means of X-ray diffraction (XRD), differential scanning calorimetry (DSC), Mössbauer spectroscopy (MS) and energy-dispersive X-rays (EDX), as-quenched and after annealing. The alloys with x≤7 suffer primary (above 550 °C) and secondary (below 680 °C) crystallisation, whereas the alloys with x≥9 only at 580 °C. Mn doping results in increase (for x≤7) and then (for x≥9) in decrease of grain size. The hyperfine field of the amorphous precursor and remainder substantially decrease with increase of Mn content x, whereas the fields in crystallites remain nearly independent of x. A paramagnetic component appears at x9 and grows with x.  相似文献   

11.
Following a brief introduction to titanium alloys and their machinability, the cutting performance of a gamma titanium aluminide intermetallic (γ-TiAl) alloy: Ti–45Al–8Nb–0.2C wt% and a burn resistant titanium (BuRTi) alloy: Ti–25V–15Cr–2Al–0.2C wt%, is compared with creep feed grinding using SiC abrasive. The work utilised 2 separate L9 Taguchi fractional factorial arrays. Typically G-ratios were a factor of 10× greater for γ-TiAl than BuRTi, with on average 10% lower maximum power and 25% lower maximum specific energy for the γ-TiAl alloy. A combination of a moderately high wheel speed: 35 m/s, low depth of cut: 1.25 mm and low feed rate: 150 mm/min, produced the lowest average workpiece surface roughness (Ra1.4 μm). Workpiece surface integrity evaluation indicated that with lower operating parameter levels, in particular a wheel speed of 15 m/s, surfaces free of burn and cracks could be produced, while at higher wheel speeds: 35 m/s, extensive workpiece surface burn was evident, with the γ-TiAl alloy suffering extensive cracking. Microhardness measurements showed in some instances slightly increased workpiece surface hardness of around 50–60HK0.025 for the BuRTi alloy and 200HK0.025 for the γ-TiAl material over respective bulk hardness values of 375HK0.025 and 400HK0.025.  相似文献   

12.
The influence of milling and subsequent annealing on the microstructural and magnetic properties of Fe90Co10 and Fe65Co35 alloys is investigated. After milling for 8 h a body-centred cubic nanostructured Fe–Co alloy forms with an average crystallite size of about 12 nm. The magnetization saturation (MS) increases 16% for Fe65Co35 and 5% for Fe90Co10 alloys by milling for 8 h. Subsequent annealing of Fe90Co10 and Fe65Co35 powders for 105 min at 550 °C improves the MS about 6 and 11%, respectively. Before annealing, the coercivity increases (up to 60 Oe) by milling for 3 h, followed by a reduction on milling for longer periods (45 h). At the initial stage of the heating, a sharp decrease in HC to 8–10 Oe occurs due to the relief of internal strain. Further heating leads to an increase in the coercivity (intermediate times) followed by a slight diminution on heating for final stage.  相似文献   

13.
The early oxidation behaviors of Mg–Y alloys (Y = 0.82, 1.09, 4.31 and 25.00 wt.%) oxidized in pure O2 have been investigated at high temperatures. The results showed that the oxidation behaviors of the Mg–Y alloys (Y = 4.31 and 25.00 wt.%) obeyed a parabolic law, while that of the Mg–Y (Y = 0.82 and 1.09 wt.%) exhibited both parabolic and linear kinetics depending on the oxidation temperature. Auger electron spectroscopy (AES), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM) and X-ray diffraction (XRD) analyses indicated that an oxide film with a single structure composed of MgO and Y2O3 had formed. Moreover, the higher the oxidation temperature was, the thicker the oxide film was. Finally, the corresponding oxidation mechanism has been discussed, and the improved oxidation resistance of the Mg–Y alloys can be due to the formation of a continuous Mg-dissolving Y2O3 protective film.  相似文献   

14.
Cd1−xCoxTe crystals (x = 0.001, 0.003, 0.005, 0.007 and 0.009) were grown by vapour phase technique. The grown diluted magnetic semiconducting (DMS) crystals were subjected to magnetization and dc-magnetic susceptibilities at room temperature. EPR spectra were recorded at 20 K for samples of all compositions. EPR spectra exhibited a broad resonance band around g  5.43. All the studies indicated the paramagnetic nature of the samples.  相似文献   

15.
The material characteristics of W2N layer and electrical properties of W/W2N/SiO2/Si metal–oxide–semiconductor (MOS) capacitors with different W2N thickness upon annealing in N2 + H2 ambient at 500 °C for 20 min are investigated. The nitrogen concentration of W2N for the W/W2N stack with thin W2N layer (≤10 nm) is lower than that for the W/W2N stack with thick W2N layer (≥15 nm). In addition, the crystallinity of W2N in the W/W2N (15 nm) stack is better than that in the W/W2N (10 nm) stack. For all capacitors, the oxide charges decrease significantly after annealing and the amount of oxide charges is independent of the W2N thickness. However, the work function (Φm) of the W/W2N (≤10 nm) stack (4.6 eV) is smaller than that of W/W2N (15 nm) stack (5.0 eV). The Φm of W/W2N (15 nm) stack is close to that of W2N single layer. After annealing, the Φm of W/W2N (15 nm) stack and W2N single layer decrease, especially for the W2N single layer. But for the W/W2N (≤10 nm) stack, the Φm increases after annealing.  相似文献   

16.
DyRh4B4 has been known to crystallize in the primitive tetragonal (pt)-structure and to exhibit a ferromagnetic transition at 12 K, the highest magnetic transition temperature in the entire series of the RRh4B4 materials [For an extensive review on the magnetic and superconducting properties on the ternary superconductors, including RRh4B4, see, for example, Ø. Fischer, in: K.H.J. Buschow, E.P. Wohlfarth (Eds.), Ferromagnetic Materials, vol. 5, Elsevier Science Publishers B.V., 1990 (Chapter 6)]. We show here that our silicon-added samples of the nominal composition DyRh4B4Si0.2 exhibit superconductivity below Tc  4.5 K and an antiferromagnetic transition below TN  2.7 K. The 12 K transition observed in the pt-DyRh4B4 is completely suppressed. Our annealed samples mainly consist of domains of the chemical composition DyRh3.9B4.2Si0.08. These domains contain two crystallographic phases belonging to the body-centred tetragonal (bct)-structure and the orthorhombic (o)-structure. We have reasons to suggest that superconductivity and antiferromagnetic ordering arise from bct-DyRh4B4 phase and, therefore, coexist below TN  2.7 K.  相似文献   

17.
This work examines the effects of extrusion parameters namely ratio and temperature on recrystallization behavior of a Mg–Zn–Zr alloy, and their consequent effects on anisotropy in the mechanical properties. Upon extrusion, the characteristic Zr-rich cores that do not recrystallize form the so-called “soft stringers” which are deformed bands elongated in the extrusion direction and curled around the extrusion axis. At higher extrusion ratio, there is more twinning contribution and the DRX response is improved, making the recryatallized grains finer and increasing the proportion of recrystallized Zr-rich cores. A basal texture after extrusion and the directional activation of tensile twinning cause anisotropy in the mechanical properties. In addition, the microstructural features such as large unrecrystallized regions and coarse crystallized grains also contribute in the strength differential effect. Further slip in the strain-hardened unrecrystallized grains is inhibited while twin activation under favorable orientation becomes easier in the coarse recrystallized grains. A higher proportion of large unrecrystallized and coarse crystallized gains in the case of lower extrusion ratio result in a much higher strength differential effect (100 MPa) in comparison to the one caused by the crystallographic texture only (25 MPa).  相似文献   

18.
The present study contemplates the application of Ru-based bimetallic alloys for hydrogen generation by hydrolysis of sodium tetrahydroborate (NaBH4). Ru and Pt, RuCu, RuPd, RuAg and RuPt (atomic ratio 1:1), PtAg, and RuxPty (atomic ratios x:y of 2:1 or 1:2), all supported over titanium oxide, were prepared. Their activity decreased in the order RuRu2Pt1 > RuPtRu1Pt2 > RuPd > RuAgPt > RuCu > PtAg. Alloying Ru with an inactive metal like Cu, Pd or Ag did not improve the performances of Ru. The catalytic ability of Ru2Pt1-TiO2 is in the range of the highest values reported so far in the literature with a hydrogen generation rate of 15.2 L(H2) min−1 g−1(RuPt). After separation from the reaction medium and rinsing with deionised water, the used Ru2Pt1-TiO2 catalyst was re-evaluated and almost the same catalytic activities as fresh catalyst were obtained during several cycles.  相似文献   

19.
We show that the variation of Tc in Er(Ni1−xPtx)2B2C (Tc  10.6 K and TN  5.7 K for x = 0) as a function of x proceeds in two steps: strong decrease of Tc for initial values of x (0 ≤ x < 0.10, Tc = 7.3 K at x = 0.1) and, thereafter, a relatively much weaker drop (almost a plateau) of Tc with further increase of x. TN exhibits a slight, almost linear, decrease over the entire range of x studied here; TN = 4.7 K for x = 0.2. Our results for x = 0.10 are in sharp disagreement with the results, namely, Tc < TN, as reported by Felner et al. [I. Felner, D. Schmitt, B. Barbara, C. Godart, E. Alleno, J. Solid State Chem. 133 (1997), 5].  相似文献   

20.
A simple model to understand the phase behavior of RNi5–H (R=La, Pr, Nd and Sm) systems is proposed based on the statistical mechanics. The essential parameters in the present model, hydrogen site energies and lattice relaxation term, were optimized so as to reproduce the experimentally determined pressure–composition isotherms (PCIs) of these four alloys. The qualitative difference in the PCI of LaNi5 from other alloys – at room temperature the former tends to show a single plateau while the latter have two plateaux – is explicable in terms of hydrogen site energies as follows. In case of LaNi5, four hydrogen sites, which are considered to belong to 3f, 12n, 6m and 12o sites, are stable in α phase, leading to a wide single plateau upto N6 where N is in RNi5HN below room temperature. At higher temperature, small differences in site energies for these sites brings the appearance of the second plateau. In case of other three alloys, on the other hand, the 6m site (or 12o) is less stable than LaNi5 in the α phase. After obtaining enough lattice relaxation energy by around N=4 through hydrogenation, the 6m site becomes favorable for hydrogen, bringing the sencond plateau spanning from N4 to N6. Predicted TC phase diagrams are also shown.  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