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.
For a set of fcc metals, our total energy calculations based on many body potentials show that activation barriers for lateral manipulation of an adatom at a step edge depend on the tip/substrate composition. Of the six homogeneous systems studied, manipulation on stepped Ag(111) showed the lowest energy barrier for adatom hopping toward the tip, although the relative probability for this process was largest on Cu(111). For a representative Cu/Pt heterogeneous system, we find lateral manipulation of a Pt adatom along a step on Pt(111) by a Cu(100) tip to be energetically much less favorable than the reverse case of a Cu adatom manipulated by a Pt(100) tip. In the case of vertical manipulation, atomic relaxations of the tip and its neighboring atoms are found to be prominent and tip-induced changes in the bonding of the adatom to its low coordinated surroundings help explain the relative ease with which an adatom next to a step edge or a kink site may be pulled as compared to that on a flat surface.  相似文献   

2.
Recent studies show that three-dimensional Ehrlich-Schwoebel (3D ES), or facet-facet, barriers of face-centered-cubic metals are substantially higher than other surface diffusion barriers. This paper presents the numerical results of 3D ES barriers for body-centered-cubic W, using classical molecular statics calculations and the nudged elastic band method. Results show that an adatom on W{110} has a diffusion barrier of 0.49 eV on the flat surface, 0.66 eV over a monolayer step, and 0.98 eV over a ridge to a neighboring {100} facet, which is one 3D ES barrier.  相似文献   

3.
We have studied the Si(001) surface with single-height steps by ab initio molecular dynamics simulations. Surface dimers were found to be unstable with respect to buckling for all geometries considered. However, the ground state reconstruction depends on the type of step. For the SA step, the c(2 × 4) geometry is induced by the step edge, while, for the SB step, the p(2 × 2) reconstruction is more stable. The binding sites and diffusion barriers for a single Si adatom were investigated via the adiabatic trajectory method. In agreement with other studies of the flat surface, fast diffusion takes place along the dimer rows. The local changes to buckling induced by the adatom are sizable and lead to changes in the activation barriers for diffusion, in particular for the path perpendicular to the dimer rows. We also investigated the diffusion of the adatom over the rebonded SB step. The calculations show that there is no additional barrier for the arrival of the adatom at the edge from the upper terrace, while a barrier of at least 1 eV exists for the arrival of the adatom from the lower edge. In step flow growth involving the rebonded SB step, most of the adatoms will thus arrive from the upper terrace.  相似文献   

4.
Zinc oxide (ZnO) nanowire synthesized from direct Zinc (Zn) vapor transport in O2 environment has been studied. The results show that the first step is the formation of ZnO film on the substrate. Then anisotropic abnormal grain growth in the form of ZnO platelets takes place. Subsequently, single-crystalline ZnO platelets grow in [0001] direction to form whiskers. During whisker growth, transformation from layer-by-layer growth to simultaneous multilayer growth occurs when the two-dimensional (2D) Ehrlich–Schwoebel (ES) barrier at the ZnO island edge is sufficiently large and the monolayer island diameter is smaller than the island spacing. As multilayered islands grows far away from the base, isotropic mass diffusion (spherical diffusion) will gradually displace anisotropic diffusion (linear diffusion), which contributes to the formation of pyramid on the top plane of the whisker. When the pyramid contains enough atomic layers, the 2D ES barrier transits to 3-dimensional ES barrier, which contributes to repeated nucleation and growth of multilayered islands or pyramids on the old pyramids. The pyramids play a critical role to taper the whisker to nanorod with a diameter less than 100 nm. The nanorod then grows to nanowire via repeated growth of epitaxial hexagonal-pyramid shape-like islands on the (0001)-plane with facets as the slope planes. During coarsening, the breakage of step motion of facets and the appearance of facets on the base of pyramids may result from the step bunching of {0001} facets, which is consistent with the existence of “2D” Ehrlich–Schwoebel barrier on the edge of (0001) facets. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

5.
Summary The free energy of surfaces along the <011> pole in gold and copper is determined to assess the effect of surface relaxation and atomic vibration on the equilibrium crystal shape of gold and copper. The Wulff construction is performed on the -plots to determine the equilibrium shape of gold and copper crystallites at different temperatures. It is shown that surface relaxation and atomic vibration do not have any discernible effect on the equilibrium shape of EAM gold or copper crystallites. The equilibrium shape of EAM gold crystallites is formed entirely from {111} and {100} facets, while that of EAM copper shows small {110} facets in addition to the {111} and {100} facets.  相似文献   

6.
Sansoz F 《Nano letters》2011,11(12):5378-5382
Surface faceting on sidewalls is ubiquitously observed during crystal growth of semiconductor nanowires. However, predicting the thermal transport characteristics of faceted nanowires relevant to thermoelectric applications remains challenging. Here, direct molecular dynamics simulations show that thermal conductivity is considerably reduced in crystalline <111> Si nanowires with periodic sawtooth faceting compared to nanowires of same size with smooth sidewalls. It is discovered that surface phonon scattering is particularly high with {100} facets, but less pronounced with {113} facets and remarkably low with {111} facets, which suggests a new means to optimize phonon dynamics for nanoscale thermoelectric devices. This anomaly is reconciled by showing that the contribution of each facet to surface phonons is due to diffuse scattering rather than to backward scattering. It is further shown that this property is not changed by addition of an amorphous shell to the crystalline core, similar to the structure of experimental nanowires.  相似文献   

7.
Epitaxy in Si technologies has to be integrated in the flow of fabrication; in most cases, it has to be selective and deposition takes place in extremely small patterns.In a first part, Si or SiGe epitaxy faceting is presented and discussed. Today, the global view is that, at first order, faceting is a kinetic phenomenon, controlled by deposition kinetic anisotropies. On the contrary, in a second part we show that highly faceted structures tend to decrease their surface energy by thermal rounding and that this phenomenon is very important when considering very small (20-40 nm) patterns.The main part reports on the combination of faceting with small size effects. Experiments consist in depositing Si0.72Ge0.28 selectively in very narrow (35-80 nm) lines oriented either along < 110 > or < 100 > crystal axis on (100) Si wafers. Preliminary observations clearly demonstrate an important {100} faceting that is often not observed or reported in literature. Our results also evaluate the lateral overgrowth of selective epitaxies. < 110 >-oriented lines lead to a certain lateral epitaxial overgrowth that is limited by {111} facets whereas epitaxies in < 100 > lines present a much larger (> 2 times) overgrowth bordered by {100} facets.Finally, we demonstrate that the edge effects have to be taken into account. Firstly, the amount of epitaxial material deposited in narrow lines depends on the line orientation, and then we propose the concept of “anisotropic loading effect”. Secondly, we found that deposition rates in small patterns are not constant with time. This corresponds to “time-nonlinear loading effects” that were never conceptualized in literature.  相似文献   

8.
Necklace-like In2O3 nanowires, which consist of In2O3 crystals (beads) and nanowires (thread), demonstrated by a simple thermal method. The shape of bead crystals based on an {111}-octahedron, of which vertices are truncated with {100} facets. According to the growth direction of the nanowires, the truncated octahedron has two different shapes; i.e., symmetric one on < 100>-nanowires and asymmetric one on < 111>-nanowires. It is shown that the growth of the bead crystals is determined by anisotropy of the growth rates of low-index facets related to their surface energy and an epitaxial relationship between the bead crystals and nanowires. Their formation mechanism and photoluminescence are discussed.  相似文献   

9.
During thin film deposition, an adatom diffuses on a flat island, down the island, or along edges of the island. Away from the flat region, the adatom generally experiences an extra migration energy – the Schwoebel-Ehrlich (ES) barrier as it has been referred as. Recently, we have proposed the concept of three-dimensional (3D) ES barrier – the facet-facet diffusion barrier. In this Letter, we propose another new concept – the step-facet barrier, with an emphasis on diffusion along the step edge. Based on these two new developments, we generalize the model of diffusion barriers near edges of islands, by classifying them as horizontal or vertical, and one to three dimensions.  相似文献   

10.
为了分析Gd吸附原子在Cu(001)、Cu(110)和Cu(111)表面的扩散机制,本文用分子动力学对该扩散过程进行模拟。模拟结果表明在Cu(001)和Cu(111)表面,Gd原子通过跳跃机制扩散;在Cu(110)表面,在[110]方向Gd原子通过跳跃机制扩散,而且多步跳跃频率很高,而在[001]方向则通过交换机制扩散。通过对扩散频率的拟合,发现在各种扩散机制都符合Arrhenius公式,从而确定了在Cu(001)和Cu(111)表面扩散势垒分别为0.19eV和0.013eV,在Cu(110)表面,沿[110]方向跳跃扩散势垒和沿[001]方向交换扩散势垒分别为0.097eV和0.33eV。  相似文献   

11.
Beginning of epitaxial growth of vapor deposition on (0 0 1) surface has been studied by use of molecular dynamics. The activation energies of motion and conversion can be calculated using an embedded atom potential at any temperatures. The activation energy of the motion of a gold adatom on (0 0 1) surface of gold is calculated to be 0.41 eV. The activation energies of dissociation in the direction of a di-adatom is 0.72 eV from the nearest neighbor to the next nearest neighbor, and 0.27 eV from the next nearest neighbor to the third neighbor. Butterfly motion of a di-adatom is 0.42+0.12 eV=0.54 eV. Tri-adatoms are classified by the angle between the two bonds in the adatoms and the lengths of the two bonds. The conversion energies of tri-adatoms have also been calculated. The activation energy for the motion of an adatom on (0 0 1) surface is much higher than those corresponding the motion on (1 1 1) surfaces.  相似文献   

12.
This study investigated the microstructural development of an electrodeposited Ni layer formed from an additive-free Watt's bath. The major texture component of the electrodeposited Ni layer is strong < 110> fiber. The electrodeposited Ni layer consists of epitaxial regions with a thickness of about 100 nm and fine columnar grains extending along the growth direction. The fine columnar grains contain a high density of twins parallel to the growth direction. These twins were formed at side {111} facets during the lateral growth of the electrodeposited Ni layer. We surmise that the twins formed in order to change the orientation so that the energy of boundary between the existing grain, on which twin related grains nucleated, could be reduced.  相似文献   

13.
Recently a controversy has developed over whether crystallographic crack growth near threshold in Ni-base superalloys occurs along {111} slip planes or {100} planes at room temperature. In this work on Nimonic API crack propagation is shown to occur on both {100} and {111} planes. The most common facet plane is {111} and this is the only orientation observed at the lowest stress intensities, but at higher stress intensities occasional {100} facets are also produced. This behaviour is compared with similar results in aluminium alloys.  相似文献   

14.
Platinum (Pt) is an outstanding catalyst for many important industrial products. Because of its high cost and scarce reserves, it is very important to improve the performance of Pt catalysts. As the metal nanocrystals (NCs) with high-index surfaces usually show very good catalytic activity because of their high density of atomic steps and kinks, the preparation of Pt NCs with high-index facets has become a very important and hot research topic recently. In this article, we report a facile synthesis of high-yield Pt NCs with a series of {hkk} high-index facets including {211} and {411} via a solvothermal method using Pt(II) acetylacetonate as the Pt source, 1-octylamine as the solvent and capping agent, and formaldehyde as an additional surface structure regulator. Multipod Pt NCs with dominant {211} side surfaces were produced without formaldehyde, while concave Pt NCs with dominant {411} surfaces formed under the influence of formaldehyde. By analyzing the products by IR spectroscopy, we found the presence of CO on the surface of concave Pt NCs with {411} surfaces prepared from the solution containing formaldehyde. It was concluded that amine mainly stabilized the monoatomic step edges, resulting in the {211} exposed surface; with addition of formaldehyde, it decomposed into CO, leading to the formation of {411} surfaces by the additional adsorption of the CO on the {100} terraces. In addition, it was found that the as-prepared Pt NCs with high-index {211} and {411} surfaces exhibited much better catalytic activity in the electro-oxidation of ethanol than a commercial Pt/C catalyst or Pt nanocubes with low-index {100} surfaces, and the catalytic activities of Pt crystal facets decreased in the sequence {411}>{211}>{100}.   相似文献   

15.
A hypomonotectic alloy of Al-4.5wt%Cd has been manufactured by melt spinning and the resulting microstructure examined by transmission electron microscopy. As-melt spun hypomonotectic Al-4.5wt%Cd consists of a homogeneous distribution of faceted 5 to 120 nm diameter cadmium particles embedded in a matrix of aluminium, formed during the monotectic solidification reaction. The cadmium particles exhibit an orientation relationship with the aluminium matrix of {111}Al//{0001}Cd and 110AlAl//11¯20> Cd, with four cadmium particle variants depending upon which of the four {111}Al planes is parallel to {0001}Cd. The cadmium particles exibit a distorted cuboctahedral shape, bounded by six curved {100}Al//{20¯23}Cd facets, six curved {111}Al/{40¯43}Cd facets and two flat {111}Al//{0001}Cd facets. The as-melt spun cadmium particle shape is metastable and the cadmium particles equilibrate during heat treatment below the cadmium melting point, becoming elongated to increase the surface area and decrease the separation of the {111}Al//{0001}Cd facets.The equilibrium cadmium particle shape and, therefore, the anisotropy of solid aluminium-solid cadmium and solid aluminium -liquid cadmium surface energies have been monitored by in situ heating in the transmission electron microscope over the temperature range between room temperature and 420 °C. The anisotropy of solid aluminium-solid cadmium surface energy is constant between room temperature and the cadmium melting point, with the {100}Al//{20¯23}Cd surface energy on average 40% greater than the {111}Al//{0001}Cd surface energy, and 10% greater than the {111}Al//{40¯43Cd surface energy. When the cadmium particles melt at temperatures above 321 °C, the {100}Al//{20¯23}Cd facets disappear and the {111}Al//{40¯43}Cd and {111}A1//{0001}Cd surface energies become equal. The {111}Al facets do not disappear when the cadmium particles melt, and the anisotropy of solid aluminium-liquid cadmium surface energy decreases gradually with increasing temperature above the cadmium melting point.The kinetics of cadmium solidification have been examined by heating and cooling experiments in a differential scanning calorimeter over a range of heating and cooling rates. Cadmium particle solidification is nucleated catalytically by the surrounding aluminium matrix on the {111}Al faceted surfaces, with an undercooling of 56 K and a contact angle of 42 °. The nucleation kinetics of cadmium particle solidification are in good agreement with the hemispherical cap model of heterogeneous nucleation.  相似文献   

16.
Density functional theory calculations are implemented in order to scrutinize indium adsorption and incorporation mechanisms in polar AlN. Indium adsorption is promoted on both polarity surfaces and adatom kinetics calculations indicate lower diffusion barriers of indium along the prismatic 〈11[`2]0 11\bar{2}0 〉 directions on (0001) as well as (000[`1] 000\bar{1} ) AlN. The latter is correlated to experimental observations of In0.24Al0.76N grown by metal organic vapour phase epitaxy, demonstrating indium concentration along the facet junctions of V-defects. This can be attributed to In surface diffusion along the 〈11[`2]0 11\bar{2}0 〉 directions of the pyramidal facets. Surface thermodynamics reveal a manifold behaviour of indium in polar AlN surfaces, significantly affected by polarity, growth stoichiometry as well as surface termination. In particular, N-rich growth conditions enhance indium incorporation on Al-terminated surfaces of both polarities, leading up to full monolayer coverage. Incorporation on N-terminated (0001) and (000[`1] 000\bar{1} ) surfaces is hindered independent of growth stoichiometry.  相似文献   

17.
Fracture experiments performed at room temperature on four test samples made of Fe-3wt.%Si single crystals with an edge crack (1 1 0)[0 0 1] (crack plane, crack front) showed approximately 45° deflections of the crack from the initial crack plane (1 1 0). This behavior appeared to be independent on loading rate. Fractographic analysis confirmed that the cracks were deviated along {1 0 0} planes and the fracture was accompanied by dislocation slip and by twinning. 3D simulations at 300 K by molecular dynamic technique in bcc iron with edge cracks of equivalent orientation indicated that the crack itself could contribute to understanding of this behavior by three processes: twinning on oblique {1 1 2} planes, which hindered growth of the original crack, and by emission of dislocations on oblique {0 1 1} and {1 2 3} planes, which led to separation of the {1 0 0} planes and might cause decohesion and subsequent cleavage fracture along the mentioned planes.  相似文献   

18.
By making a step on one surface ( ) of a rectangular small paralellepiped copper crystal, dislocations could be created by the molecular dynamic method. The dislocation created was not a complete edge dislocation but a pair of Heidenreich-Shockley partial dislocations. Each time a dislocation was created, the stress on the surface was released. Small copper crystals having a notch were pulled (until fracture), compressed and buckled by use of the molecular dynamic method. An embedded atom potential was used to represent the interaction between atoms. Dislocations were created near the tip of the notch. A very sharp yield stress was observed. The results of high speed deformations of pure silicon small crystals using the molecular dynamics are presented. The results suggest that plastic deformation may be possible for the silicon with a high speed deformation even at room temperature. Another small size single crystal, the same size and the same surfaces, was compressed using molecular dynamic method. The surfaces are {110}, {112} and {111}. The compressed direction was [111]. It was found that silicon crystals are possible to be compressed with a high speed deformation. This may suggest that silicon may be plastically deformed with high speed deformation.  相似文献   

19.
Surface atoms have fewer interatomic bonds than those in the bulk that they often relax and reconstruct on extended two-dimensional surfaces. Far less is known about the surface structures of nanocrystals. Here, we show that coherent diffraction patterns recorded from individual nanocrystals are very sensitive to the atomic structure of nanocrystal surfaces. Nanocrystals of Au of 3-5 nm in diameter were studied by examining diffraction intensity oscillations around the Bragg peaks. Both results obtained from modelling the experimental data and molecular dynamics simulations strongly suggest inhomogeneous relaxations, involving large out-of-plane bond length contractions for the edge atoms (approximately 0.2 A); a significant contraction (approximately 0.13 A) for {100} surface atoms; and a much smaller contraction (approximately 0.05 A) for atoms in the middle of the {111} facets. These results denote a coordination/facet dependence that markedly differentiates the structural dynamics of nanocrystals from bulk crystalline surfaces.  相似文献   

20.
Nanocrystalline Au films were in situ strained in a high resolution transmission electron microscope, which demonstrated that the diffusion-assisted intergranular fracture was the dominant failure mode. Grain orientation with respect to grain boundaries (GBs) imposes important effect on the crack propagation and blunting. The low surface energy and high diffusion mobility of {111} planes lead to a notch-like crack. The stress concentration at the tip may help breaking {111} planes layer by layer and thus advance the crack. Cracks can be diverted from the preset path by GBs and grow into the grain interior, which has never been revealed by other experiments and molecular dynamics simulations.  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