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.
InAs nanowires were grown on GaAs substrates by the Au-assisted vapour–liquid–solid (VLS) method in a gas source molecular beam epitaxy (GSMBE) system. Passivation of the InAs nanowires using InP shells proved difficult due to the tendency for the formation of axial rather than core–shell structures. To circumvent this issue, AlxIn1?xAs or AlxIn1?xP shells with nominal Al composition fraction of x=0.20, 0.36, or 0.53 were grown by direct vapour–solid deposition on the sidewalls of the InAs nanowires. Characterisation by transmission electron microscopy revealed that the addition of Al in the shell resulted in a remarkable transition from the VLS to the vapour–solid growth mode with uniform shell thickness along the nanowire length. Possible mechanisms for this transition include reduced adatom diffusion, a phase change of the Au seed particle, and surfactant effects. The InAs–AlInP core-shell nanowires exhibited misfit dislocations, while the InAs–AlInAs nanowires with lower strain appeared to be free of dislocations.  相似文献   

2.
The crystal growth of 3C-SiC onto silicon substrate by Vapor–Liquid–Solid (VLS) transport, where a SiGe liquid phase is fed with propane, has been investigated. Three sample configurations were used. In a preliminary approach, the VLS growth of SiC was conducted directly onto Si substrate using a Ge film as liquid catalyst. It led to the growth of a thick continuous SiC polycrystalline layer which was floating over a SiGe alloy located between the silicon substrate and the topping SiC layer. In the second configuration, a thin seeding layer of 3C-SiC grown by chemical vapor deposition (CVD) was used and the VLS growth was localized using a SiO2 mask. The liquid phase was a CVD deposited SiGe alloy. The growth of a few hundred nanometers thick 3C-SiC epitaxial layer was demonstrated but the process was apparently affected by the presence of the oxide which was dramatically etched at the end. In the last configuration, the silicon substrate was patterned down to 10 μm and a thin seeding layer of 3C-SiC was grown by CVD onto this patterned substrate. The liquid phase was again a CVD deposited SiGe alloy. In this last configuration, the presence of epitaxial SiC was evidenced but it grew as trapezoidal islands instead of an uniform layer.  相似文献   

3.
We report a simple, versatile, two-step fabrication technique for synthesizing a core–shell nanorod array whose architecture is specifically suited for use as an electrode in a dye sensitized solar cell (DSSC). The particular structure fabricated by us consists of a parallel array of 5 μm long and 150–200 nm wide Ag nanorod cores, each coated with a 15–20 nm thick ZnO shell. Importantly, the shell thickness is roughly uniform throughout the length of the rods, which are free standing but distinctly separated from each other. This would allow the dye to penetrate freely and cover the ZnO surface completely in a DSSC.  相似文献   

4.
We prepared SiO2@Ag core–shell nanospheres: silver nanoparticles (~4 ± 2 nm in diameter) coated silica nanospheres (~50 ± 10 nm in diameter). The preparation route is a modification of the Stöber method, and involves the preparation of homogeneous silica spheres at room temperature, combined with the deposition of silver nanoparticles from Ag+ in solution, by using water/ethanol mixtures, tetraethyl-orthosilicate as Si source and silver nitrate as Ag source in a single-pot wet chemical route without an added coupling agent or surface modification, which leads to the formation of core@shell homogeneous nanospheres. We present the preparation and characterization of the SiO2@Ag core–shell nanospheres and also of bare silica spheres in the absence of silver, and propose a reaction mechanism for the formation of the core–shell structure.  相似文献   

5.
Dielectrics, polarizing optical microscopic and electro-optical measurements have been carried out on a core/shell quantum dot Cd1?xZnxS/ZnS dispersed ferroelectric liquid crystal (FLC). In the present study, quantum dots were dispersed into two different concentrations of 0.1 and 0.25 wt./wt.% in pure FLC. The electro-optical parameters of pure and QDs dispersed FLC were carried out as a function of applied voltage. A significant improvement in optical response time of QDs dispersed FLC system is one of the major finding of the present study which may be useful for fabrication of faster liquid crystal system.  相似文献   

6.
Liquid phase epitaxial growth of III–V solid solutions invariably involves contact between the multi-component saturated or undersaturated liquid and solid phases which are not in thermodynamic equilibrium because either the number of components is different or the composition of the layer to be grown differs from the composition of the underlying layer. The non-equilibrium system must relax to the final equilibrium state through some intermediate ones.The main point of the present review is to show that all non-equilibrium systems encounted in hetero-LPE come through the following stages of relaxation:
  • 1.1. Partial dissolution of the solid with simultaneous formation of a thin diffusive dividing layer (0.5–3 nm thick) (DDL) at the solid/liquid interface (in the subsurface region of the solid). The layer contains all the components of the given system and in some cases the quasi-equilibrium between the saturaed multi-component liquid and the solid diffusive dividing layer can be observed experimentally. If the DDL is mismatched to the substrate the former must be strained. So, the Gibbs potential of the solid increased additionally and the liquid must become supersaturated by additionally dissolving the substrate.
  • 2.2. The nucleation and the growth of centres of a new phase at some points on the solid/liquid interface with the simultaneous dissolution of the solid at other areas of the same interface (mechanism of “etch-back and regrowth”).
  • 3.3. The formation of a continuous epitaxial dividing layer (EDL), separating a bulk solid from a multi-component liquid. At this stage the system stratifies into a three-layer configuration: multi-component liquid/EDL/substrate, further relaxation is limited by a solid state diffusion.
It is only during dedicated experiments on selected systems that stages 1 to 3 can be observed separately in their more or less pure form. Usually, the first and the second stages almost escape experimental detection due to an enormous variability in time scales and an inherent randomness of events during the “fast” stage 2. So, in practice the varied combination of relaxation modes is observed. It is, however, true that, among many parameters which influence the mode of non-equilibrium solid/liquid interface relaxation, the most critical are the temperature of isothermal contact between liquid and solid phases and the lattice mismatch between a substrate and a new solid phase. This concerns not only the magnitude of the mismatch but its sign as well.  相似文献   

7.
Epitaxial single-crystal ZnS/Si core–shell nanowires have been synthesized via a two-step thermal evaporation method. The epitaxial growth is due to the close match of crystal structure between zinc blende ZnS and diamond-like cubic Si. The nanowires have a uniform diameter of 80–200 nm and a length of several to several tens of micrometers. Single-crystal Si nanotubes can be obtained by chemical etching of the ZnS/Si core–shell structure. Characteristics of field-effect transistors (FETs) fabricated from the Si nanotubes suggests that the Si tubes show weak n-type semiconductivity with a mobility of about 3.7×10?2 cm2/(V s), which is 1 order larger than that of intrinsic Si.  相似文献   

8.
The grain boundary groove shapes for equilibrated solid neopentylglycol (NPG) solution (NPG–3 mol% D-camphor) in equilibrium with the NPG–DC eutectic liquid (NPG–36.1 mol% D-camphor) have been directly observed using a horizontal linear temperature gradient apparatus. From the observed grain boundary groove shapes, the Gibbs–Thomson coefficient (Г), solid–liquid interfacial energy (σSL) of NPG solid solution have been determined to be (7.5±0.7)×10?8 K m and (8.1±1.2)×10?3 J m?2, respectively. The Gibbs–Thomson coefficient versus TmΩ1/3, where Ω is the volume per atom was also plotted by linear regression for some organic transparent materials and the average value of coefficient (τ) for nonmetallic materials was obtained to be 0.32 from graph of the Gibbs–Thomson coefficient versus TmΩ1/3. The grain boundary energy of solid NPG solution phase has been determined to be (14.6±2.3)×10?3 J m?2 from the observed grain boundary groove shapes. The ratio of thermal conductivity of equilibrated eutectic liquid to thermal conductivity of solid NPG solution was also measured to be 0.80.  相似文献   

9.
The principle of interferometers and its applicability to our research on crystal growth can be understood through assembling interferometers. In particular, practical skills such as techniques for assembling interferometers and selecting optical components, which are not covered by general textbooks, can be learned.  相似文献   

10.
Phase-field simulations are used to examine tip velocity and shape selection in free dendritic growth of a pure substance into an undercooled melt in the presence of a density change between the solid and liquid. The dendrite is assumed to grow two-dimensionally inside a Hele-Shaw cell. The phase-field model is coupled with a previously developed two-phase diffuse interface model to simulate the flow in the liquid that is induced by the density change. The predicted dependence of the dendrite tip growth Péclet number on the relative density change is compared with an available analytical solution and good agreement is obtained. The simulations verify that the dendrite tip selection parameter, modified to account for the different densities of the solid and liquid phases, is independent of the relative density change.  相似文献   

11.
Nanosized iron core and barium titanate shell microstructure was generated within a silicate glass of composition 23.1 Na2O, 23.1 BaO, 23.0 TiO2, 7.6 B2O3, 5.8 Fe2O3, 17.4 SiO2 by first reducing it at 893 K for ½ h and then subjecting it to heat treatment at 759 K for 4 h. Transmission electron microscopy showed the composite particles to have a mean diameter of 3.9 nm. The nanocomposite exhibited both ferroelectric and ferromagnetic behavior. The dielectric constant peak was not prominent because of a small thickness of the barium titanate phase. The magnetic hysteresis loop showed an asymmetric behavior giving rise to a small exchange bias field. This is believed to arise due to exchange interaction between the ferromagnetic iron core and the thin layer of Fe3O4 on the core surface with a spin glass-like behavior. The magnetization under zero-field cooled (ZFC) and field cooled (FC) conditions indicated superparamagnetic behavior at temperatures higher than 300 K. The optical absorption spectra exhibited a peak at around 325 nm. This was analyzed satisfactorily on the basis of a metal core–oxide shell nanoconfiguration. The extracted values of metal core conductivity showed a metal insulator transition for iron core diameters less than 2.4 nm. The present synthesis approach will lead to newer multiferroic nanocomposites and glasses with multifunctionalities.  相似文献   

12.
Zn powders were milled by stainless steel balls in the aqueous solution whose PH value were adjusted to 3, 7, and 11, respectively. Zn(OH)2 and ZnO powders could be synthesized via this novel solid–liquid ball milling method. The compositions and the microstructures of the as-milled products were characterized by X-ray diffraction and SEM, respectively. The experimental results showed that the higher PH value is helpful to accelerate the reaction between Zn and H2O. ZnO is the preferential product in the neutral and alkaline solutions. As a preferential product, a large number of ZnO could be obtained in the neutral and alkaline solutions after reaction. Finally, the phase transformation regularities of the as-milled products were also discussed.  相似文献   

13.
Tin oxide (SnO2) nanorods were synthesized through an aqueous hexamethylenetetramine (HMTA) assisted synthesis route and their structural evolution from core–shell type faceted pyramidal assembly was investigated. Structural analysis revealed that the as-synthesized faceted SnO2 structures were made of randomly arranged nanocrystals with diameter of 2–5 nm. The shell thickness (0–80 nm) was dependent on the molar concentration of HMTA (1–10 mM) in aqueous solution. It was revealed that the self-assembly was possible only with tin (II) chloride solution as precursor and not with tin (IV) chloride solution. At longer synthesis hours, the pyramidal nanostructures were gradually disintegrated into single crystalline nanorods with diameter of about 5–10 nm and length of about 100–200 nm. The SnO2 nanorods showed high sensitivity towards acetone, but they were relatively less sensitive to methane, butane, sulfur dioxide, carbon monoxide and carbon dioxide. Possible mechanisms for the growth and sensing properties of the nanostructures were discussed.  相似文献   

14.
Measurements of the metastable zone and solubility for flunixin meglumine–ethanol system were obtained. The solubility was measured within the temperature range from 288.15 to 328.15 K. The mole fraction solubility was correlated satisfactorily with the temperature by the equation: xeq=2.35×10?12e0.07121T. The value of enthalpy of dissolution, enthalpy of fusion and enthalpy of mixing were determined to be 49.04, 64.03 and ?14.99 kJ mol?1 respectively. The metastable zone width of flunixin meglumine was measured by an electric conductivity method. A comparison of the nucleation temperatures from electric conductivity measurement and from focused beam reflectance measurement (FBRM) shows that both detection techniques give almost the same results for flunixin meglumine. The nucleation parameters of flunixin meglumine in ethanol were determined from the metastable zone data. Over the equilibrium temperature range from 312.28 to 325.55 K, the nucleation rate constant was varied from 0.00001 to 0.00120 #/m2 min, whereas the nucleation order was varied from 2.23022 to 3.39299. The obtained high values of nucleation order indicated a high rate of nucleation.  相似文献   

15.
Transport of ionic species in the melt and their partitioning at the solid–liquid interface during growth of lithium niobate was studied under the influence of intrinsic and external electric fields. A Mn-doped lithium niobate (Mn:LiNbO3) single crystal was grown via the micro-pulling-down (μ-PD) method with electric current injection at the interface. Mn ions were accumulated or depleted at the interface, depending on the sign of the injected current. The electric current injection induced an interface electric field as well as a Coulomb force between the interface and Mn ions. The electric field modified the transportation of Mn ions and their partitioning into the crystal, while the Coulomb force led to adsorption or rejection of Mn ions at the interface in addition to Mn concentration change due to the electric field. Effect of the Coulomb force was often observed to be larger on Mn concentration at the interface than that of the induced electric field, and dominated the redistribution of Mn in the solid. It has been experimentally and analytically shown that Mn concentration partitioned into the crystal can be obtained by multiplying Mn concentration at the interface by a field-modified partition coefficient, kE0, instead of the conventional equilibrium partition coefficient, k0.  相似文献   

16.
Viscosity of Bi100?xZnx (x = 0, 8, 16, 50, 80, 85, 89, 100) liquid alloys has been measured by means of the oscillating crucible method. Temperature dependences of the viscosity coefficient have revealed anomalous behavior in the vicinity of the demixing curve. The maximum on background of Arrhenius type dependence of viscosity is most resolved for melt of critical concentration. The viscosity measurements results are analyzed from the point of view of concentration fluctuations. It is concluded that concentration–concentration fluctuations are responsible for the anomalous behavior of viscosity.  相似文献   

17.
The fluctuation effects at the boundary between the internal solid and external liquid regions of the Earth’s core are considered within the vitrification model. If the internal core is characterized by a relatively small static shear modulus and shear acoustic oscillations, this phase transition becomes slightly stepwise and can be accompanied by critical phenomena. The corresponding fluctuation corrections to the thermodynamic derivatives are calculated with arbitrary critical indices. The propagation and absorption of seismic waves, which become weakly anisotropic due to the anisotropic “liquid” inclusions (which arise in the hysteretic boundary region due to the Earth’s rotation and viscous stress effect), are considered. The resulting estimates are compared with the existing geophysical data.  相似文献   

18.
A model of oriented growth of (0001) ZnO whiskers on sapphire substrates via the vapor–crystal mechanism using the catalytic properties of gold islands is proposed. The morphological transition from the primary pyramidal ZnO structures to hexagonal ZnO whiskers is described in terms of the minimization of the free energy density of three-dimensional heteroepitaxial islands.  相似文献   

19.
We examined the volume and viscosity of Zr–Cu–Al glass-forming liquid alloys to clarify the origin of a frozen free volume in glassy alloys. Since an excess free volume imparts toughness and ductility to glassy alloys, we attempted to increase this volume in glass structures so that they could be used as engineering materials. The maximum frozen excess free volume was observed in the ternary eutectic composition of the Zr–Cu–Al alloy system; however, its origin remains unclear. We attempted to reveal the mechanism of the formation of the frozen excess free volume in Zr–Cu–Al glassy alloys.  相似文献   

20.
The possibility of AlN growth using Li–Al–N solvent was investigated. Based on theoretical prediction, we selected Li3N as a suitable nitrogen source for AlN growth. First, vapor phase epitaxy using Li3N and Al as source materials was performed to confirm the following reaction on the growth surface: Li3N+Al=AlN+3Li. The results suggest that the reaction proceeds to form AlN on the substrate under appropriate conditions. Next, AlN growth using Li–Al–N solvent was carried out. The Li–Al–N solvent was prepared by annealing of mixtures composed of Li3N and Al. The results imply that AlN was formed under an Al-rich condition. Moreover, it was found that Li was swept out from AlN grains during growth. The results suggest that AlN growth using Li–Al–N solvent might be a key technology to obtain an AlN crystal boule.  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