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 375 毫秒
1.
The structure, phase composition, and thermal stability of the industrial zirconium alloys, namely, E110 (Zr–1% Nb) and E635 (Zr–1% Nb–0.3% Fe–1.2% Sn), which are subjected to high‐pressure torsion (HPT) at room temperature (RT), 200, and 400 °С have been studied. HPT of Zr‐alloys at RT (10 revolutions) leads to the formation of grain–subgrain nano‐sized structure and to increase the microhardness by 2.1…2.8 times. The increase in the HPT temperature to 200–400 °С leads to the increase in the structural‐element average size. The structural‐element size in the complexly alloyed E635 alloy in all cases is lower compared with the E110 alloy. The hardening of the alloys after HPT at RT and 200 °С is close, and at 400 °С is much less. HPT initiates the α‐Zr → (ω‐Zr + β‐Zr) transformation, which is the main factor for alloys hardening. The α‐Zr → (ω‐Zr + β‐Zr) transformation in the E635 alloy occurs less quickly. The maximum amount (ω‐Zr + β‐Zr) phase in the structure of the alloys is observed after HPT at RT and 200 °C, and the minimum ? at 400 °C. During heating, the alloys undergo the reverse (ω‐Zr + β‐Zr) → α transformation which depends on both the alloy composition and HPT temperature.
  相似文献   

2.
Ti–Nb–Ta–Zr system alloys are receiving more attention for biomedical material component applications. However, the deformation behavior of the Ti–Nb–Ta–Zr system has not been evaluated to date. Therefore, the deformation behavior of Ti–Nb–Ta–Zr alloys with different Nb contents was investigated in this study.The behaviors of loading–unloading stress–strain curves of Ti–20Nb–10Ta–5Zr and Ti–25Nb–10Ta–5Zr air-cooled after final heating of the manufacturing process are similar to that obtained in metastable β type titanium alloys that have the shape memory effect. Therefore, the shape memory effect was expected in Ti–20Nb–10Ta–5Zr and Ti–25Nb–10Ta–5Zr alloys. The elastic deformation of Ti–30Nb–10Ta–5Zr disobeyed Hooke's law. However, stress or strain-induced martensite (SIM) is not observed on the loading–unloading stress–strain curve. The deformation mechanism of Ti–25Nb–10Ta–5Zr changes with varying its microstructure. In Ti–25Nb–10Ta–5Zr air-cooled after final heating, the microstructure consisted of an ω phase in a β phase. The stress for inducing martensite in a β phase, σM, was nearly equal to the yielding stress, σy. Therefore, stress-induced martensitic transformation and movement of dislocations occurred together. In Ti–25Nb–10Ta–5Zr water-quenched after final heating of the manufacturing process, the microstructure consisted of a single β phase, where σM is lower than σy. Therefore, stress-induced martensitic transformation occurred before yielding.  相似文献   

3.
We report the results of ab initio molecular dynamics simulation for the structural transition of Zr during two distinct quenching processes (Q1: 4.3 × 1013 K/s, Q2: 2.0 × 1014 K/s). In both the quenching processes, structural transition details have been analyzed by pair correlation functions g(r) and bond pair analysis technique. It is shown that the liquid Zr transforms to a metastable bcc phase (β-Zr) at the temperature about 1000 K as quenched at the rate of 4.3 × 1013 K/s. When quenched at 2.0 × 1014 K/s, however, the crystallization is suppressed and the liquid Zr is frozen into a glass state. The bond pair analysis reveals that the dominant bond pairs in the liquid and glass states are the 1551, 1541, 1431, 1661 and 1441, indicating that the short range order in both states mostly consists of icosahedral, tetrahedral and bcc clusters.  相似文献   

4.
In this work, the effects of Nb content on microstructure and corrosion behaviors of biomedical Zr–Nb alloys were systematically studied. The results of XRD analysis and optical microscopy indicated that the experimental Zr–Nb alloys had a duplex structure of α and β phases, and the content of β phase increased with the increase of Nb content. The electrochemical impedance spectroscopy (EIS) studies showed an improvement on the resistance of the spontaneous oxide film with increasing Nb content. The EIS data, fitted by Rs(QpRp) model, suggested a single passive film formed on the experimental material surfaces. Polarization tests in Hank's solution revealed a nobler electrochemical behavior of the Zr–Nb alloys after alloying Nb to pure Zr. The corrosion resistance increased with increasing Nb content, as indicated by lower corrosion current densities and passive current densities and higher pitting potentials. The major components on the surfaces of the corroded Zr–Nb alloy samples detected by XPS were ZrO2 and Nb2O5. The biocompatibility of Zr–Nb alloys was primarily evaluated by culturing L-929 cells in the extraction media of Zr–Nb alloy samples and excellent results were obtained. All of these above results suggested that the Zr–22Nb alloy, among the experimental alloys, showed a promising potential for biomedical applications.  相似文献   

5.
Abstract

Phase transformations and mechanical properties of both Ti–29Nb–13Ta–4·6Zr and Ti–39Nb–13Ta–4·6Zr (wt–%) alloys were investigated. The microstructure of the 29Nb alloy is sensitive to solution and aging treatment. Ice water quenching from the solution treatment temperature resulted in (β+α") microstructure but air or furnace cooling led to a mixture of (β+ω). The formation of the orthorhombic α" martensite thus suppresses ω formation in the ice water quenched 29Nb alloy. Cooling rate from the solution treatment temperature also has a significant effect on the formation of α and ω phases during subsequent isothermal aging below the ω start temperature: slow cooling enhances ω but depresses α formation. This cooling rate dependence of aged microstructure was attributed to α" martensite acting as precursor of the α phase, thus providing a low energy path to the precipitation of a at the expense of ω. Phase transformation in the 39Nb alloy is more sluggish than that in the 29Nb alloy, owing to the presence of the higher content of β stabiliser Nb. For the 29Nb alloy, Young's modulus and mechanical properties are sensitive to the fraction of phases, and change significantly during aging, in contrast with the 39Nb alloy.  相似文献   

6.
Analytical expressions are given for the phase boundaries of diagrams of state for Zr-Nb and Zr-Sn systems. Thermodynamically justified temperature dependences of heat capacity of Zr-1% Nb, Zr-2.5% Nb, Zr-1.5% Sn, and Zr-2.5% Sn alloys in two-phase regions are given and recommended for use. The heats of second-order phase transformations α(Zr) ? β(Zr) in alloys of Zr-Nb and Zr-Sn binary systems are obtained, and the heats of monotectoid reaction α(Zr)+β(Nb) ? β(Zr) in the Zr-Nb system and of peritectoid reaction β(Zr)+Zr4Sn ? α(Zr) in the Zr-Sn system are calculated.  相似文献   

7.
Ti–Nb–Fe–Sn alloys with relatively low Nb content, located near the phase boundary of (β + ω)/β, are designed on the basis of electron-to-atom (e/a) ratio, d-electron alloy design concept, and Mo equivalent (Moeq) aiming at low Young's modulus comparable to human bone. The effect of Sn content and Nb content on the microstructure and the mechanical properties is investigated in Ti–5Nb–3Fe–(0–6)Sn (at%) and Ti–(3–9)Nb–3Fe–4Sn (at%) alloys. The composition dependence of Young's modulus and tensile strength of Ti–Nb–Fe–Sn alloys is analyzed in terms of the phase stability, ω phase, and recrystallization texture. Both Nb and Sn are effective in suppressing the athermal ω phase and stabilizing the β phase. The recrystallization texture is strongly influenced by the content of Sn and Nb. A strong {110}β<001>β Goss texture is formed in the Ti–5Nb–3Fe–(2–4)Sn and Ti–(3–5)Nb–3Fe–4Sn alloys. The Ti–5Nb–3Fe–4Sn alloy exhibits an exceptionally low Young's modulus of 30 GPa due to the combined effects of low stability of the β phase, a small amount of ω phase, and a strong Goss texture.  相似文献   

8.
Fatigue crack growth (fcg) behavior of cold-worked and stress relieved Zr–2.5 Nb was studied in the longitudinal (with and without hydrides) and transverse direction at ambient temperature and load ratio of 0.1 using compact tension samples. Fatigue loading in the transverse direction (distribution of both hard and soft grains) showed facet formation on the fracture surface and the highest ΔKth whereas loading in the longitudinal direction (distribution of primarily soft grains) showed no facet formation and a lower ΔKth. Hydrided Zr–2.5 Nb loaded in the transverse direction showed large facets with the lowest ΔKth. Texture influenced fcg at low ΔK but not at higher ΔK.  相似文献   

9.
The hot deformation behavior, dynamic recrystallization, and texture evolution of Ti–22Al–25Nb alloy in the temperature range of 950–1050 °C and strain rate range of 0.001–1 s?1 is investigated by plane‐strain compression testing on the Gleeble‐3500 thermo‐mechanical simulator. The results show that the flow stress decreases with the increase of temperature and decrease of strain rate. Besides, the flow curves appear a serrate oscillation at a strain rate of 0.1 s?1 for all the temperature ranges, which may result from instability such as flow localization or micro‐cracking. The flow behavior can be expressed by the conventional hyperbolic sine constitutive equation and the calculated deformation activation energy Q in the (α2 + B2) and B2 regions are 631.367 and 304.812 kJ mol?1, respectively. The microstructure evolution is strongly dependent on the deformation parameters, and dynamic recrystallization (DRX) is the dominant softening mechanism in the (α2 + B2) region, including discontinuous dynamic recrystallization (DDRX), and continuous dynamic recrystallization (CDRX). In addition, the ηbcc‐fiber of {110} <001> is the dominant texture component in deformed Ti–22Al–25Nb alloy. It is observed that the weakening of the deformation texture is accompanied by the occurrence of DRX, which can be attributed to the large misorientation between DRX grains and neighboring B2 matrix induced by the rotation of DRX grains toward the preferred slip systems.
  相似文献   

10.
We have studied the relaxor behavior of sol-gel derived Ba(Zr x Ti1− x )O3 (0.30≤ x≤0.70) thin films. The plausible mechanism of the relaxor behavior has been analyzed from the dielectric data and micro-Raman spectra. Substitution of Zr+4 for Ti+4 in BaTiO3 lattice reduces its long-range polarization order yielding a diffused paraelectric to ferroelectric phase transition. The solid solution system is visualized as a mixture of Ti+4 rich polar region and Zr+4 rich regions and with the increase in Zr contents the volume fraction of the polar regions are progressively reduced. At about 25.0 at% Zr contents the polar regions exhibit typical relaxor behavior. The degree of relaxation increases with Zr content and maximizes at 40.0 at% Zr doped film. The frequency dependence of the polar regions follows Vogel-Fulcher relation with a characteristic cooperative freezing at freezing temperature (T f). Below Tf, a long range polarization ordering was ascertained from the polarization hysteresis measurement.  相似文献   

11.
Ti–XNb–10Ta–5Zr (mass %) alloys based on nominal compositions of Ti–35Nb–10Ta–5Zr, Ti–30Nb–10Ta–5Zr, and Ti–25Nb–10Ta–5Zr were fabricated through powder metallurgy and forging and swaging processes for biomedical applications. The tensile deformation mechanisms of the Ti–25Nb–10Ta–5Zr, Ti–30Nb–10Ta–5Zr, and Ti–35Nb–10Ta–5Zr alloys were investigated in situ by X-ray diffraction analysis under several loading conditions.Under the loading conditions, the X-ray diffraction peaks of all the specimens shifted to higher angles than those obtained under the unloading conditions. For the Ti–30Nb–10Ta–5Zr alloy, the elastic deformation is considered to progress continuously in a different crystal direction although after the elastic strain reaches elastic limit in the crystal direction where the elastic limit is the smallest, slip deformation occurs in that crystal direction. The elastic modulus of this alloy appears to decrease in terms of strain over the proportional limit. Thus, the elastic deformation behavior of the Ti–30Nb–10Ta–5Zr alloy does not obey Hooke's law.  相似文献   

12.
In this study, we investigated the effect of Ta content on the mechanical properties of Ti–30Nb–XTa–5Zr fabricated by a powder metallurgy method, for biomedical applications. The Ta content ranged from 0% to 20 mass%.The microstructures of Ti–30Nb–XTa–5Zr that contain less than 5 mass% Ta comprise β phase and an ω phase. The tensile properties of Ti–30Nb–XTa–5Zr change with a change in their deformation mechanisms. The deformation mechanisms of Ti–30Nb–XTa–5Zr, which contains less than 10 mass% Ta, is the stress-induced martensite (SIM) transformation, while that of Ti–30Nb–XTa–5Zr, which contains over 20 mass% Ta, is the slip mechanism. The minimum elastic modulus is obtained in Ti–30Nb–10Ta–5Zr, which comprises a single β phase.  相似文献   

13.
Nickel-free Ti–22Nb–6Zr alloys were fabricated by conventional powder metallurgy sintering method. X-ray diffractometer (XRD) investigation showed that the as-sintered alloys mainly consisted of β phase, with a few needle-like α phase precipitates. Differential scanning calorimetry (DSC) measurement in the temperature ranging from −70 °C to 400 °C and constant stress thermal cycling test by dynamic mechanical analysis (DMA) were unable to reveal the martensitic start temperature of sintered Ti–22Nb–6Zr alloys. Therefore low temperature compression tests were carried out to evaluate their phase transformation behavior indirectly. There was an obvious drop of both Young’s modulus and recoverable strain at −85 °C  −80 °C in the Young’s modulus-temperature and recoverable strain–temperature curves of sintered Ti–22Nb–6Zr alloys respectively, which was attributed to the occurrence of thermal elastic martensitic transformation at this temperature. At the testing temperature of −85 °C, a superelasticity of as high as 5.9% was achieved in the sintered alloys. The results had revealed that sintered Ti–22Nb–6Zr alloys own a great superelasticity intrinsically and would exhibit a much greater superelasticity at room temperature if their martensitic transformation start temperature (Ms) were closer to room temperature. Along with their noble biocompatibility, sintered nickel free Ti–22Nb–6Zr alloys are thus thought to be potentially competitive biomaterials for biomedical applications.  相似文献   

14.
Abstract

Metal release from implantable metals and the properties of oxide films formed on alloy surfaces were analyzed, focusing on the highly biocompatible Ti–15Zr–4Nb–4Ta alloy. The thickness and electrical resistance (Rp) of the oxide film on such an alloy were compared with those of other implantable metals. The quantity of metal released during a 1-week immersion test was considerably smaller for the Ti–15Zr–4Nb–4Ta than the Ti–6Al–4V alloy. The potential (E10) indicating a current density of 10 μA cm?2 estimated from the anodic polarization curve was significantly higher for the Ti–15Zr–4Nb–4Ta than the Ti–6Al–4V alloy and other metals. Moreover, the oxide film (4–7 nm thickness) formed on the Ti–15Zr–4Nb–4Ta surface is electrochemically robust. The oxide film mainly consisted of TiO2 with small amounts of ZrO2, Nb2O5 and Ta2O5 that made the film electrochemically stable. The Rp of Ti–15Zr–4Nb–4Ta was higher than that of Ti–6Al–4V, i.e. 0.9 Ω cm2 in 0.9% NaCl and 1.3 Ω cm2 in Eagle's medium. This Rp was approximately five-fold higher than that of stainless steel, which has a history of more than 40 years of clinical use in the human body. Ti–15Zr–4Nb–4Ta is a potential implant material for long-term clinical use. Moreover, E10 and Rp were found to be useful parameters for assessing biological safety.  相似文献   

15.
Abstract

The transformation of precipitated phases of Zr50·5Cu34·5?xNi4Al11Agx alloy master ingots with Ag substitution of Cu was studied in detail by phase analysis. The precipitated (Zr–Cu) rich phases deteriorate the glass forming ability (GFA) of Zr50·5Cu34·5Ni4Al11. Two new (Zr–Cu) rich phases, A1 with bcc structure and a?=?0·339 nm and A2 with fcc superlattice structure and a?=?1·21 nm, were identified by a transmission electron microscope. When x?=?2, A1 phase disappears, and A2 phase remains and is suppressed gradually with further Ag addition. When x?=?13, one precipitated phase with Ag more than 13 at-% begins to deteriorate GFA. In the view of chemistry, the precipitation of (Zr–Cu) rich phases means that the interaction between Cu and Zr atoms is rather drastic. The addition of Ag weakens the interaction between Cu and Zr. The similar competition mechanism proposed by the authors plays an important role in suppressing precipitated phases and improving GFAs.  相似文献   

16.
We present an ab initio calculation of interlayer magnetic coupling for Fe/Nb multilayers using the self-consistent full-potential linearized augmented-plane-wave (FLAPW) method. For this calculation, we have constructed supercells consisting of bcc Fe and Nb multilayers in Fe/Nb/Fe sandwich geometry stacked along (001) direction. In the supercells two Fe layers are separated by Nb layers ranging from 1 to 11 layers. We have calculated the total energy of the system as a function of Nb spacer layer thickness. For each spacer layer thickness, we have done three calculations corresponding to para, ferro and antiferromagnetic ordering of Fe atoms. The interlayer magnetic coupling is obtained from the energy difference of the systems in which Fe layers are antiferromagnetically and ferromagnetically ordered. We find that the interlayer magnetic coupling oscillates with increasing Nb spacer thickness in agreement with the experimental results. The induced magnetic moment is also found to be oscillating with increasing Nb spacer layer thickness.  相似文献   

17.
The recent progress in the development of nanostructured composites is described for Zr‐base multicomponent alloys as a typical example for such materials. These advanced composite materials are attractive candidates for structural as well as functional applications. The combination of high strength with high elastic strain of fully nanocrystalline and glassy alloys renders them quite unique in comparison to conventional (micro‐)crystalline materials. However, one major drawback for their use in engineering applications is the often limited macroscopic plastic deformability, despite the fact that some of these alloys show perfectly elastic‐plastic deformation behavior. To improve the room temperature ductility of either fully nanocrystalline or amorphous alloys, the concept of developing a heterogeneous microstructure combining a glassy or nanostructured matrix with second‐phase particles with a different length‐scale, has recently been employed. This review describes the composition dependent metastable phase formation in the Zr‐(Ti/Nb)‐Cu‐Ni‐Al alloy system, which in turn alters the mechanical properties of the alloys. We emphasize the possibilities to manipulate such composite microstructures in favor of either strength or ductility, or a combination of both, and also discuss the acquired ability to synthesize such in‐situ high‐strength composite microstructures in bulk form through inexpensive processing routes.  相似文献   

18.
This article provides an overview of metastable β titanium alloys either being utilized or being considered for use in orthopedic applications. The effects of thermomechanical processing on the mechanical properties (e.g., elastic modulus, tensile, wear and high cycle fatigue performance) of Ti‐15Mo‐0.2O, Ti‐12Mo‐6Zr‐2Fe (TMZF), Ti‐29Nb‐13Ta‐4.6Zr and Ti‐35Nb‐7Zr‐5Ta are reviewed. The osteointegration behavior of Ti‐29Nb‐13Ta‐4.6Zr and Ti‐35Nb‐7Zr‐5Ta‐O alloys is also presented.  相似文献   

19.
Corrosion behavior of a newly developed multifunctional β-type Ti–23Nb–0.7Ta–2Zr–O (mol%, TNTZO) alloy in Ringer's physiological solution was evaluated by open circuit potential, potentiodynamic polarization and X-ray photoelectron spectroscopy (XPS) techniques. Corrosion property of Ti–6Al–4V was also measured for comparison. The results showed that the TNTZO alloy possesses much better corrosion property than the Ti–6Al–4V alloy, corroborated by a high corrosion potential and broad passive region, which is attributed to the stable and inert passive TiO2 film modified by the oxides of Nb, Ta and Zr on the surface of the TNTZO alloy.  相似文献   

20.
Nanostructured Ti–31.0Fe–9.0Sn and sub-micrometer structured Ti–39.3Nb–13.3Zr–10.7Ta (wt.%) β-type alloys, exhibiting different microstructures and dissimilar mechanical properties, have been prepared by copper mold casting. The microstructure, mechanical behavior and corrosion resistance, in simulated body fluid, of both alloys have been investigated and compared to those of commercial Ti–6Al–4V. Nanoindentation experiments reveal that the Ti–31.0Fe–9.0Sn rods exhibit very large hardness (H  9 GPa) and high Young's modulus. Conversely, the Ti–39.3Nb–13.3Zr–10.7Ta alloy is mechanically softer but it is interesting for biomedical application because of its rather low Young's modulus (E  71 GPa). Concerning the corrosion performance, Ti–35Nb–7Zr–5Ta shows a corrosion behavior comparable to Ti–Al6–V4, with no potential breakdown up to 0.4 V vs. Ag|AgCl. On the contrary, the Ti–31.0Fe–9.0Sn alloy exhibits a more anodic corrosion potential, but the value is still less negative than for pure elemental Fe and Ti. From all these properties and because of the absence of toxic elements in the compositions, the Ti–39.3Nb–13.3Zr–10.7Ta and Ti–31.0Fe–9.0Sn alloys are attractive for use as metallic biomaterials.  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