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 390 毫秒
1.
The GaN thin film deposited on an amorphous glass substrate was analyzed by using transmission electron microscopy with a new automated crystal orientation mapping tool. Film deposition was made at 600°C for 4 h by the hyperthermal neutral beam (HNB) source. Columnar crystals oriented to the [0001] direction without significant disordering were clearly observed. Electron diffraction patterns indicated that the crystals have mainly two different zone axes, [2 $\bar 1$ $\bar 1$ 0] and [10 $\bar 1$ 0]. This crystallographic and microstructural information provides the guidance for future works for the HNB source to obtain GaN thin films of higher quality on amorphous substrates.  相似文献   

2.
X-ray diffraction and electron microscopy were used to study the development of the γ-? martensitic transformation (MT) upon tensile deformation of single crystals of (I) the Fe-17% Cr-12% Ni-2% Mn-0.75% Si and (II) Fe-18% Cr-12% Ni-2% Mo-0.015% C (wt %) austenitic stainless steels as a function of the crystal-axis orientation and the test temperature T. It has been shown that a decrease in the test temperature to T<173 K in single crystals of steels I and II with a low stacking fault energy (γ0=0.01–0.015 J/m2) leads to a γ-?-α’ MTs upon plastic deformation. It has been established that the degree of deformation preceding the γ-? MT depends on the crystal-axis orientation and the γ0 magnitude. In the [011] and $[\bar 1 11]$ crystals, the γ-? MT upon tension is developed already at early stages of plastic flow, at ?≤3%, whereas in the $[\bar 1 23]$ and [012] crystals it occurs after a substantial deformation by slip, at ?=16–70%. In the [001] crystals, no γ-? MT is revealed by X-ray diffraction, but 1–2% ? phase is observed by electron microscopy. The physical cause for the observed orientation dependence in the γ-? MT is related to the effect of external stresses σ on the degree of splitting of perfect dislocations a/2〈110〉 to Shockley partial dislocations a/6〈211〉, which form nuclei of the ? phase.  相似文献   

3.
Electron microscopy has been used to study orientation relationships (ORs) and stable misorientations between corresponding directions of the icosahedral ι phase and the β phase (CsCl type structure) of the solid solution in quenched two-phase Al61Cu26Fe13 alloys with a dendritic structure. Calculated values of the azimuthal misorientations have been obtained for the directions of the reciprocal lattices of the ι and β phases in diffraction patterns A5 l ,[113]β; A2 l , [111]A5β for the basic $OR[110]_\beta \left\| {A5_l , [\bar 11\bar 1]_\beta } \right\|A2_l $ and close orientation conditions such as $[110]_\beta \left\| {A5_l , [\bar 110]_\beta } \right\|A2_l ; [11\bar 1]_\beta \left\| {A3_l , [\bar 110]_\beta } \right\|A2_l ; [111]_\beta \left\| {A2_l , [\bar 110]_\beta } \right\|A2_l $ . Two possible mechanisms of stabilization of the azimuthal misorientations revealed are discussed; first, the realization of the $OR[111]_\beta \left\| {A2_l , [\bar 110]_\beta } \right\|A2_l $ in quenched Al-Cu-Fe alloys and, second, the stability of the arising misorientations as a result of the formation of a layer with a modulated icosahedral structure at the boundaries between the β and gi phases.  相似文献   

4.
A transmission electron microscopy investigation on the phase decomposition of B2-ordered (Ni,Co)Al supersaturated with Ni and Co has revealed the precipitation of (Ni,Co)2Al which has not been expected from the reported equilibrium phase diagram. The (Ni,Co)2Al phase has a hexagonal structure and takes a rodlike shape with the long axis of the rod parallel to the 〈111〉 directions of the B2 matrix. By aging at temperatures below 873 K, a long period superlattice structure appears in the hexagonal (Ni,Co)2Al phase. The orientation relationship between the (Ni,Co)2Al precipitates and the B2-(Ni,Co)Al matrix is found to be (0001)p//(111)B2 and [ $ \bar 1 $ 2 $ \bar 1 $ 0]p//[ $ \bar 1 $ 10]B2, where the suffix p and B2 denote the (Ni,Co)2Al precipitate and the B2-(Ni,Co)Al matrix, respectively. (Ni,Co)Al hardens appreciably by the fine precipitation of the (Ni,Co)2Al phase.  相似文献   

5.
After the primary recrystallization of a preliminarily deformed (110)[001] single crystal, the texture also has the preferred (110)[001] orientation. Furthermore, the texture contains weak orientations, a major part of which is formed at the sample surface and can be described by a spectrum of scattered orientations {120}〈210〉…{351}〈103〉. A further heating leads to two concurrent processes taking place in the samples, i.e., the normal growth of Goss grains and secondary recrystallization. Abnormally grown crystals are represented by a quartet of orientations related with the initial Goss orientation by a rotation around [011], [01 $\bar 1$ ], [101], and [10 $\bar 1$ ] axes at an angle of ~30°. The crystallographic relationship between the initial and final grain orientations can be explained by their closeness to special misorientations as follows: Σ9, Σ19a, Σ27a, and Σ33a (rotation around 〈110〉 axes to close angles).  相似文献   

6.
The microstructure of stress-induced martensite(SIM) in the nanocrystalline NiTi alloy was investigated by means of transmission electron microscopy(TEM). The result shows that the multi-variant structure of the martensite is suppressed and only single-variant martensitic twins form after tensile deformation when the grain size is smaller than80 nm. The normal directions of the(001)B190twin planes are all within the range of 45° from the axial direction of the wire. The angle between twin crystals(1"11)Mand(111)Tof the SIM is also found to be smaller than that of thermally induced martensite in nanocrystalline NiTi.  相似文献   

7.
The decomposition and precipitation behaviors of a quenched Cu–15wt%Sn alloy as a function of aging temperature were investigated using transmission electron microscopy (TEM). Focused ion beam (FIB) was employed to assist TEM specimen preparation. At 300 °C, the decomposition of the supersaturated α′ phase occurred at grain boundaries, displaying a cellular morphology. The lamellae were found with ζ and α phases, rather than with the equilibrium ε and α phases. The ζ and α phases exhibit a well-defined orientation relationship (OR) as $ (1\bar{1}0)_{\alpha } //(0001)_{\zeta } ,\;[11\bar{2}]_{\alpha } //[\bar{1}2\bar{1}0]_{\zeta } $ . On the other hand, at 320 °C, only incipient lamellar structures of several micron meters were observed, which were composed of the δ and α phases. At the same time, abundant intragranular precipitation of the ε phase in the form of platelets was observed, and OR as $ (1\bar{1}1)_{\alpha } //(001)_{\varepsilon } , $ [110] α //[100] ε exists between ε phase and the α phase. These contrasting precipitation behaviors are discussed from the viewpoint of crystallographic coherency of these phases.  相似文献   

8.
The dependence of critical shear stresses for twinning τ cr tw on the stacking-fault energy γ0, the nitrogen concentration C N (wt %), and the test temperature has been studied for [ $ \bar 1 $ 11] single crystals of austenitic stainless steels upon tensile deformation. It is shown that τ cr tw ~ γ0/b 1 in nitrogen-free steels, while it exhibits a nonmonotonic dependence on the nitrogen concentration C N in nitrogen-alloyed steels. An increase in τ cr tw at C N = 0.3 wt % and its decrease at C N ≥ 0.5–0.7 wt % are determined by the competition of two contributions to τ cr tw , namely, a decrease in γ0 and a solid-solution hardening with increasing nitrogen concentration.  相似文献   

9.
In order to investigate $\left\{ {10\bar 12} \right\}$ tensile twinning behavior, the magnesium single crystal was deformed by compressing along the $\left[ {2\bar 1 \bar 10} \right]$ direction at room temperature, as $\left\{ {10\bar 12} \right\}$ tensile twinning easily takes place when the compression direction is perpendicular to the c-axis. Numerous $\left\{ {10\bar 12} \right\}$ primary tensile twins were activated during deformation, and the Schmid factor (SF) criterion was applied to the six $\left\{ {10\bar 12} \right\}$ twin variants. The analysis shows that the majority of the $\left\{ {10\bar 12} \right\}$ primary twins belong to high SF variants, and high SF twin boundaries provided nucleation sites for low SF variants. The $\left\{ {10\bar 12} \right\}$ secondary tensile twins were formed inside the high SF of wide $\left\{ {10\bar 12} \right\}$ primary twin bands, and the basal plane of the $\left\{ {10\bar 12} \right\}$ secondary twin was tilted about 60° with respect to the original parent matrix. In the case of the $\left\{ {10\bar 12} \right\}$ secondary tensile twin, relatively low SF variants were activated while counterparts with higher SF variants were absent.  相似文献   

10.
The studies were performed on D3 tool steel hardened after austenitizing at 1050 °C during 30 min and tempering at 200-700 °C. Based on the diffraction studies performed from the extraction replicas, using electron microscopy, it was found that after 120-min tempering in the consecutive temperatures, the following types of carbides occur: $$ 200\;^\circ {\text{C}} \to \upvarepsilon + \upchi + {\text{ Fe}}_{ 3} {\text{C}},\quad 3 50\;^\circ {\text{C}} \to \upvarepsilon + \upchi + {\text{ Fe}}_{ 3} {\text{C,}} $$ $$ 500\;^\circ {\text{C}} \to \upchi + {\text{ M}}_{ 3} {\text{C }} + {\text{ M}}_{ 7} {\text{C}}_{ 3} ,\quad 600\;^\circ {\text{C}} \to \upchi + {\text{ M}}_{ 3} {\text{C }} + {\text{ M}}_{ 7} {\text{C}}_{ 3} , $$ $$ 700\;^\circ {\text{C}} \to {\text{M}}_{ 3} {\text{C }} + {\text{ M}}_{ 7} {\text{C}}_{ 3} . $$ Apart from higher mentioned carbides, there are also big primary carbides and fine secondary M7C3 carbides occurring, which did not dissolve during austenitizing.  相似文献   

11.
The development of the γ-α’ martensitic transformation (MT) upon tensile deformation of single crystals of austenitic stainless steels of compositions (wt %) Fe-17% Cr-12% Ni-2% Mn-0.75% Si (I) and Fe-18% Cr-12% Ni-2% Mo-0.015% C (II) has been studied as a function of the crystal-axis orientation and test temperature by X-ray diffraction and electron microscopy. It has been established that the orientation dependence of the slip deformation preceding the γ-α’ MT is determined by two factors, namely, the orientation dependence of slip deformation preceding the γ-? MT and the orientation dependence of the work U necessary for the formation of the α’-martensite crystals. The orientation dependence of slip deformation preceding the γ-? MT leads to the γ-α’ MT in the [\(\bar 1\)11], [\(\bar 1\)23], [011], and [012] crystals with different defect densities and, correspondingly, at different stress levels. In the [001] crystals, no γ-α’ MT is observed macroscopically since the γ-? MT in these crystals is suppressed. It has been established that the γ-α’ MT in the [\(\bar 1\)11], [011], [\(\bar 1\)23], and [012] crystals can be developed at T = 300 K after preliminary deformation at T = 77 K. The development of the γ-α’ MT at T = 300 K is physically related to the growth of the α’-martensite nuclei formed upon plastic deformation at T = 77 K.  相似文献   

12.
The last decade has seen a remarkable surge in x-ray characterization methods (Willmott, An Introduction to Synchrotron Radiation, John Wiley & Sons, Inc., New York, 2011). Imaging with x-rays has evolved from simple radiography, to image internal structure and diagnose injury, to a full-fledged tool for nanoscale characterization (Holt et al., Annu Rev Mater Res 43:1, 2013). Central to this development has been the advent of high-brilliance synchrotron and free electron laser sources of x-rays. The high degree of spacial coherence of the resulting beams has enabled novel imaging methods. Of these, coherent diffraction imaging has proven highly successful at imaging the structure in nano materials (Miao et al., Nature 400:342, 1999). In addition, this imaging method can be combined with Bragg diffraction to image strain with high sensitivity (Pfeifer et al., Nature 442:63, 2006; Robinson and Harder, Nat Mater 8:291, 2009).  相似文献   

13.
Phase equilibria in the quasi-ternary system Ag2Se-CdSe-Ga2Se3 were investigated by differential thermal and x-ray phase analysis methods. Phase diagrams of nine vertical sections were constructed. The boundaries of seven single-phase fields were determined which are solid solution ranges of system components and intermediate phases. We constructed the isothermal section at 820 K and the liquidus surface projection, and have determined the position in the system of six invariant processes with the participation of liquid: $ {\text{L}}_{{{\text{U}}_{1} }} + {\upzeta} {\leftrightarrows} {\upbeta} + {\upeta} $ L U 1 + ζ ? β + η (1145 K), $ {\text{L}}_{{{\text{U}}_{ 2} }} + \upzeta \leftrightarrows \upgamma + \upeta $ L U 2 + ζ ? γ + η (1138 K), $ \text{L}_{{U_{3} }} + \upeta \leftrightarrows \updelta + \upgamma $ L U 3 + η ? δ + γ (1113 K), $ {\text{L}}_{{{\text{E}}_{ 1} }} \leftrightarrows \upbeta + \updelta + \upeta $ L E 1 ? β + δ + η (1083 K), $ {\text{L}}_{{{\text{E}}_{ 2} }} \leftrightarrows \upalpha + \upbeta + \upvarepsilon $ L E 2 ? α + β + ε (969 K), $ {\text{L}}_{{{\text{E}}_{ 3} }} \leftrightarrows \upbeta + {\updelta} + \upvarepsilon $ L E 3 ? β + δ + ε (963 K). Two invariant processes in the sub-solidus part, $ \upbeta + \updelta \leftrightarrows \upeta + \uplambda $ β + δ ? η + λ and $ \upbeta + \updelta \leftrightarrows \upvarepsilon + \uplambda $ β + δ ? ε + λ at 968 and 938 K, respectively, were investigated as well.  相似文献   

14.
In support of the selection of structural materials for heat exchangers in helium-cooled high temperature reactors, the oxidation behaviour of the Ni-base chromia-former alloy 230 was investigated at 850 °C in diluted helium atmosphere with a low water vapour content. In such a media, the equivalent partial pressure of oxygen (imposed by the $ P_{{{\text{H}}_{2} {\text{O}}}} $ / $ P_{{{\text{H}}_{2}}} $ ratio) is very low ( $ P_{{{\text{O}}_{ 2} }}^{\text{eq}} $ around 10?16 Pa). The equivalent partial pressure of oxygen has no straight influence on the parabolic rate constant (k p); on the other hand, $ P_{{{\text{H}}_{2} }} $ and $ P_{{{\text{H}}_{2} {\text{O}}}} $ demonstrate a complex influence on k p. Photoelectrochemistry analyses revealed that this oxide could simultaneously contain two types of cationic defects. Specific oxidation tests with D2O showed that the oxide scale also contains hydrogen. A mechanist model is proposed in order to describe the scale growth using both cationic defects. Those theoretical results show, at least qualitatively, how $ P_{{{\text{H}}_{2} }} $ and $ P_{{{\text{H}}_{2} {\text{O}}}} $ may concurrently influence the oxidation rate.  相似文献   

15.
High purity iron and a low carbon, low silicon steel were oxidised at temperatures of 800–1,200 °C, in atmospheres of N2–H2–H2O and N2–O2–H2O. Scales of wüstite grew at low oxygen potentials, and of FeO/Fe3O4/Fe2O3 at high oxygen potentials, both according to parabolic kinetics after an initial transient period. The iron and steel behaved similarly in the O2/H2O gases, but not in H2/H2O, where the steel oxidised much more slowly than the iron. The rate for steel increased with $ p_{{H_{2} O}} $ at fixed $ p_{{O_{2} }} , $ but for iron was almost independent of $ p_{{H_{2} O}} , $ whilst rates for both metals increased with $ p_{{O_{2} }} $ at fixed $ p_{{H_{2} O}} $ . These results are discussed using point defect models involving hydroxyl anions and cation vacancies. Scaling rates in O2/H2O also increased with $ p_{{H_{2} O}} , $ a result attributed to gas phase transport within oxide pores which were present in the scales, but absent in wüstite grown in H2/H2O.  相似文献   

16.
Titanium alloys are extensively used in a variety of applications because of their good mechanical properties, high biocompatibility, and corrosion resistance. Recently, ??-type Ti alloys containing Ta and Nb have received much attention because they feature not only high specific strength but also biocorrosion resistance, no allergic problems, and biocompatibility. A Ti-25Ta-25Nb ??-type titanium alloy was subjected to severe plastic deformation (SPD) processing by accumulative roll bonding and investigated with the aim to observe the texture developed during SPD processing. Texture data expressed by pole figures, inverse pole figures, and orientation distribution functions for the (110), (200), and (211) ??-Ti peaks were obtained by XRD investigations. The results showed that it is possible to obtain high-intensity share texture modes ({001}??110??) and well-developed ?? and ??-fibers; the most important fiber is the ??-fiber ({001} $ \left\langle {1\bar{1}0} \right\rangle $ to {114} $ \left\langle {1\bar{1}0} \right\rangle $ to {112} $ \left\langle {1\bar{1}0} \right\rangle $ ). High-intensity texture along certain crystallographic directions represents a way to obtain materials with high anisotropic properties.  相似文献   

17.
The transient oxidation stage of a model metal alloy thin film was characterized with in situ ultra-high vacuum (UHV) transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS) and analytical high-resolution TEM. We observed the formations of nanosized NiO and Cu2O islands when Cu-5at.%Ni(001) was exposed to oxygen partial pressure, $ {\text{pO}}_{ 2} = 1 \times 10^{ - 4} \,{\text{Torr}} $ and various temperatures in situ. At 350 °C epitaxial Cu2O islands formed initially and then NiO islands appeared on the surface of the Cu2O island, whereas at 550 °C NiO appeared first. XPS and TEM revealed a sequential formation of NiO and then Cu2O islands at 550 °C. The temperature-dependent oxide selection may be due to an increase of the diffusivity of Ni in Cu with increasing temperature.  相似文献   

18.
In this work, the thermal stability of Cu-In alloys with 45.0 and 41.2 at.% In nominal compositions was investigated by differential scanning calorimetry (DSC), scanning electron microscopy, wavelength dispersive spectroscopy, and in-situ synchrotron x-ray powder diffraction (S-PXRD) over a temperature range from 25 up to 400 °C. The studied samples are mainly composed of a Cu11In9 phase together with minor amounts of the B phase (based on the NiAs-Ni2In type structure) and, in one of the samples, with a minor amount of pure In. No evidence of the Cu10In7 (41.2 at.% In) phase was detected, not even in the sample with 41.2 at.% In nominal overall composition. The combined use of the S-PXRD and DSC techniques allowed us to identify two phase transitions involving the Cu11In9 phase, one of them corresponding to the $ \upeta^{\prime} \rightleftharpoons {\text{B}} + {\text{Cu}}_{11} {\text{In}}_{9} $ reaction at T = 290 °C and the other to the peritectic $ \upeta^{\prime} + {\text{L}} \rightleftharpoons {\text{Cu}}_{11} {\text{In}}_{9} $ reaction at T = 311 °C.  相似文献   

19.
The corrosion behavior of magnesium single crystals with various crystallographic orientations was examined in this study. To identify the effects of surface orientation on the corrosion behavior in a systematic manner, single-crystal specimens with ten different rotation angles of the plane normal from the [0001] direction to the $ [ 10\overline{1} 0] $ direction at intervals of 10° were prepared and subjected to potentiodynamic polarization and potentiostatic tests as well as electrochemical impedance spectroscopy (EIS) measurements in 3.5?wt.% NaCl solution. Potentiodynamic polarization results showed that the pitting potential (E pit) first decreased from ?1.57?V SCE to ?1.64?V SCE with an increase in the rotation angle from 0° to 40°, and then increased to ?1.60?V SCE with a further increase in the rotation angle to 90°. The results obtained from potentiostatic tests are also in agreement with the trend in potentiodynamic polarization tests as a function of rotation angle. A similar trend was also observed for the depressed semicircle and the total resistances in the EIS measurements due to the facile formation of MgO and Mg(OH)2 passive films on the magnesium surface. In addition, the amount of chloride in the passive film was found first to increase with an increase in rotation angle from 0° to 40°, then decrease with a further increase in rotation angle, indicating that the tendency to form a more protective passive film increased for rotation angle near 0° [the (0001) plane] or 90° [the $ ( 10\overline{1} 0) $ plane].  相似文献   

20.
The adsorption of ionic surfactants on the water/air or water/hydrocarbon interface is considered. One form of the well-known Gibbs equation takes into account the surface excess of the amphiphilic ion in the compact layer, or monolayer, Γ 2 m (2D adsorption), and the differential of the electrical potential of this layer. This expression is modified using some simplifying assumptions. The dependence of the surface tension, σ, on the activity of the amphiphilic ion, a 2, degree of gegen-ions binding in the compact layer, β, and Γ 2 m is transformed into the following relationship: $- \frac{{d\sigma }} {{RTd\ln a_2 }} = \Gamma _2^m \left\{ {2 - (1 - \beta )\frac{{d\ln \Gamma _2^m }} {{d\ln a_2 }}\left[ {\frac{1} {{\left[ {1 - \left( {\Gamma _2^m /\Gamma _2^{m\infty } } \right)} \right]^{1 + t} }} - \frac{{2b}} {{RT}}\Gamma _2^m } \right]} \right\}.$ Here Γ 2 m denotes the Γ 2 m value at complete filling of the adlayer, t = ?1, 0, or +1 for the two-phase model of partition, for immobile or mobile monolayer respectively, b is the cohesion constant; both the long-tailed ion and the gegen-ion are single-charged. The usefulness of the proposed equation is discussed.  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