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 250 毫秒
1.
The technique of selective photo-etching known for GaAs was successfully applied to (001)GaAs0.6P0.4 and it can also be used for the characterization of phosphorus-rich GaAs1?xPx layers and for GaP. Using this technique misfit dislocations in GaAs1?xPx up to now mainly examined with the transmission electron microscope (TEM) in the GaAs1?xPx transition layer could be examined in a density range which is hardly accessible for observation with the TEM. Misfit dislocations are also observed in GaP and GaAs1?xPx layers when these are heavily doped with nitrogen. The dislocation density of GaAs1?xPx layers of constant composition is compared with predictions from a model of Abrahams et al. for the dislocation density of GaAs1?xPx as a function of the rate of grading. Within this model our results can only be explained when the asymmetry factor m between the length to distance ratio of misfit dislocations in the transition layer is about two orders of magnitude greater than the value previously derived from TEM studies of GaAs1?xPx layers.  相似文献   

2.
The misfit dislocation configurations in InAs islands as well as in more or less continuous layers grown on (001) oriented GaAs substrates were studied by weak-beam and high-resolution electron microscopy. The islands are confined by {101} and {111} facets where the aspect ratio (height/lateral extension) can be affected by the growth conditions. It is possible to grow well-defined islands as well as relatively continuous layers by MOVPE under As-stabilized conditions. At constant deposition parameters the growth is characterized by islands of different sizes (but with constant aspect ratio) in various strain states depending on their dislocation content. Coherently strained islands without any dislocation can be observed for heights up to 23 ML InAs, or otherwise, up to a maximal island extension of about 12 nm (for the particular aspect ratio ≈︂0.585). With further increase of island height and lateral extension, the introduction of dislocations becomes favourable. Independent of the island size, the layer thickness and the dislocation density, a residual elastic strain of about εr = —0.8% remains after relaxation. This means, about 88% of the total misfit strain of ε = —6.686 × 10—2 were compensated by Lomer dislocations. These sessile Lomer dislocations lie in the island interior only, where single 60° dislocations were observed exclusively in their near-edge regions. With increasing island size and/or layer thickness some close-spaced 60° dislocations occur additionally within the interfacial region. The Lomer dislocations that are always located 4 monolayers (ML) above the InAs/GaAs interfacial plane result from the well-known fusion of two 60° slip dislocations. These 60° dislocations have been nucleated 7 … 8 ML above the interface at surface steps on the {111} facets confining the islands. Based on our experimental observations a new mechanism is proposed that explains the origin of these 60° dislocations. Their further fusion to sessile Lomer dislocations that compensate the misfit strain most efficiently occurs in the way as commonly accepted.  相似文献   

3.
Substrates and epitactic layers of GaP are etched under optimized conditions with the defect-revealing, preferential R-C etchant and subsequently examined using high-magnification (about 2500X) optical microscopy techniques. It is possible then to distinguish two new phenomena, viz. etch pits characteristic of dislocation loops in the substrate, and inclined dislocation dipoles in the layer. For layers grown on dislocation-free substrates we find that (i) the surface densities of both defects are equal (~5 × 105cm-2), and (ii) the average diameter of the dislocation loops in the substrate is roughly the same as the average distance between the two dislocations of the dislocation dipoles (0.5?1μm). Hence the perfection of these layers is determined by interfacial dislocation loops. Because the density of dislocation loops is only about a factor of two lower in highly (~1 × 105 cm-2) dislocated substrates, growth on these substrates results in layers which have even slightly lower dislocation density than layers grown on dislocation-free substrates. In the former case also single dislocations in the substrate propagate into the layer. Minority-carrier lifetime data indicate that minority-carrier recombination at dislocations is a restrictive factor for the luminescence quality of layers grown both on dislocation-free and highly-dislocated substrates.  相似文献   

4.
Transmission electron microscopy (TEM) studies of defects in AlxGa1?xN layers with various Al mole fractions (x=0.2, 0.4) and polarities were carried out. The samples were grown by ammonia molecular beam epitaxy on sapphire substrates and consisted of low-temperature AlN (LT-AlN) and high-temperature AlN (HT-AlN) buffer layers, a complex AlN/AlGaN superlattice (SL) and an AlxGa1?xN layer (x=0.2, 0.4). It was observed that at the first growth stages a very high density of dislocations is introduced in both Al-polar and N-polar structures. Then, at the interface of the LT-AlN and HT-AlN layers half-loops are formed and the dislocation density considerably decreases in Al-polar structures, whereas in the N-polar structures such a behavior was not observed.The AlN/AlGaN superlattice efficiently promotes the bend and annihilation of threading dislocations and respectively the decrease of the dislocation density in the upper AlxGa1?xN layer with both polarities.The lattice relaxation of metal-polar Al0.2Ga0.8N was observed, while N-polar Al0.2Ga0.8N did not relax. The dislocation densities in the N-polar Al0.2Ga0.8N and Al0.4Ga0.6N layers were 5.5×109 cm?2 and 9×109 cm?2, respectively, and in metal-polar Al0.2Ga0.8N and Al0.4Ga0.6N layers these were 1×1010 cm?2 and 6×109 cm?2, respectively.Moreover, from TEM images the presence of inversion domains (IDs) in N-polar structures has been observed. The widths of IDs varied from 10 to 30 nm. Some of the IDs widen during the growth of the AlN buffer layers. The IDs formed hills on the surface of the N-polar structures.  相似文献   

5.
Dislocation structure of GexSi1?x films (x=0.4?0.8) grown by molecular-beam epitaxy on Si(001) substrates was studied by means of transmission electron microscopy. It was found that the density of edge MDs formed at the early stage of plastic strain relaxation in the films could exceed the density of 60° MDs. In our previous publications, a predominant mechanism underlying the early formation of edge misfit dislocations (MD) in GexSi1?x/Si films with x>0.4 was identified; this mechanism involves the following processes. A 60° glissile MD provokes nucleation of a complementary 60° MD gliding on a mirror-like tilted plane (111). A new edge MD forms as a result of interaction of the two complementary 60° MDs, and the length of the newly formed edge MD can then be increased following the motion of the “arms” of the complementary 60° MDs. Based on this scenario of the edge MD generation process, we have calculated the critical thickness of insertion of an edge MD into GeSi layers of different compositions using the force balance model. The obtained values were found to be more than twice lower than the similar values for 60° MDs. This result suggests that a promising strategy towards obtaining dislocation arrays dominated by 90° dislocations in MBE-grown GexSi1?x/Si films can be implemented through preliminary growth on the substrate of a thin, slightly relaxed buffer layer with 60° MDs present in this layer. The dislocated buffer layer, acting as a source of threading dislocations, promotes the strain relaxation in the main growing film through nucleation of edge MDs in the film/buffer interface. It was shown that in the presence of threading dislocations penetrating from the relaxed buffer into the film nucleation of edge MDs in the stressed film can be initiated even if the film thickness remains small in comparison with the critical thickness for insertion of 60° MDs. Examples of such unusual MD generation processes are found in the literature.  相似文献   

6.
Single- and double-heterojunction structures of AlxGa1?xP/GaP were grown by liquid phase epitaxy and their opto-electronic properties were characterized. Using a starting growth temperature of 900°C, a background impurity level in undoped Al0.5Ga0.5P of 1 × 1016 cm-3 (p-type) is obtained. The growth rate of AlxGa1?xP was very low. An almost intrinsic layer exists at the p-n AlxGa1?xP/GaP heterojunction interface due to the interdiffusion of the dopants. By using a rotating slide liquid phase epitaxial growth technique, we have made a high efficiency light modulator from a AlxGa1?xP/GaP double heterostructure. The employment of a dummy crystal resulted in very good reproducibility of the GaP waveguide layer thickness. Light is well confined in the GaP layer and a voltage difference, Vπ, of 6 V is enough to obtain the phase difference, π, between TE and TM modes.  相似文献   

7.
A solution is presented suitable for revealing lattice defects on (1 1 1 ) phosphorous faces of GaP and InP. It is possible to distinguish between pits originated by grown-in dislocations and microdefects such as perfect loops, faulted loops and precipitates and/or inclusions. Moreover it is also possible to reveal large stacking faults of Shockley type and microtwin lamellae. One to one correlations have been given by means of transmission electron microscopy.  相似文献   

8.
The influence of the metamorphic buffer design and epitaxial growth conditions on the electrical and structural characteristics of metamorphic In0.38Al0.62As/In0.37Ga0.63As/In0.38Al0.62As high electron mobility transistor (MHEMT) nanoheterostructures has been investigated. The samples were grown on GaAs(100) substrates by molecular beam epitaxy. The active regions of the nanoheterostructures are identical, while the metamorphic buffer In x Al1 ? x As is formed with a linear or stepwise (by Δ x = 0.05) increase in the indium content over depth. It is found that MHEMT nanoheterostructures with a step metamorphic buffer have fewer defects and possess higher values of two-dimensional electron gas mobility at T = 77 K. The structures of the active region and metamorphic buffer have been thoroughly studied by transmission electron microscopy. It is shown that the relaxation of metamorphic buffer in the heterostructures under consideration is accompanied by the formation of structural defects of the following types: dislocations, microtwins, stacking faults, and wurtzite phase inclusions several nanometers in size.  相似文献   

9.
The structural defects and microhardness of Bi2Te3‐xSex whiskers (x = 0, 0.2 and 0.4 at% Se) grown by physical vapour deposition (PVD) method have been investigated. Concentric pairs of dislocation loops were observed on the as‐grown surfaces of short hexagonal prisms. A systematic study of dislocations in these crystals was carried out by chemical etching technique. The effects of Se doping, annealing and quenching on the mechanical properties have also been studied on the prism faces of Bi2Te3‐xSex whiskers. (© 2008 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

10.
本文研究了InP/GaP晶格失配界面的电特性。HRTEM图象表明在界面存在90°位错缺陷的应变缓释。ECV表明界面存在高密度载流子层。AFM图象表明本研究中获得了粗糙度为2.48nm的良好InP异质外延层。并对于InP界面给出了一个基于费米能级钉扎的模型来解释观察到的电性质。  相似文献   

11.
Orthorhombic TmMnO3 (o-TMO) thin films have been epitaxially stabilized on (110) SrTiO3 substrates by pulsed laser deposition (PLD) technique. The microstructure and strain relaxation mechanism of o-TMO thin films are analyzed using transmission electron microscopy. It is shown that major defects in the films are misfit dislocations with Burgers vectors of type ap〈010〉 and ap〈110〉, whereas ap〈110〉 dislocations tend to dissociate into partial dislocations with Burgers vectors of type 1/2ap〈110〉. Strain in o-TMO films is relaxed by misfit dislocations as well as surface fluctuations, which is different from most of the previous studies of the perovskite thin films.  相似文献   

12.
Measurements of positron lifetime τ and of the shape parameter S of the Doppler-broadened annihilation line are used to study bulk and defect properties in GaP and GaAs1−xPx. τ and S decrease linearly with the composition x of GaAs1−xPx layers. A second lifetime component (τ2 = 290 ps) observed in as-grown GaP is attributed to stoichiometric P vacancies. After neutron irradiation of GaP and GaAs0.13P0.87 positrons are trapped by Ga vacancies (τ2 = 250 ps). These vacancies anneal out in two stages at 200–550 °C and 550–800 °C.  相似文献   

13.
Using electron microscopy methods it has been determined that there are numerous crystalline inclusions in evaporated layers of amorphous selenium. The density, dimensions, modification and distribution in the cross section of a layer are essentially dependent upon the method of layer preparation. The maximum micro-inclusion density is observed in layers evaporated at a substrate temperature (ts) between 8 to 30°C. These layers also possess crystallites of monoclinic modification. In layers evaporated at ts ? 50°C inclusions of trigonal modification prevail.The parameters of the deep trapping levels for holes and electrons have been determined from photodischarge kinetics. The origin of these levels is found to be related to the phase transition. The hole traps are caused by the crystalline formation of trigonal modification, and the electron traps by crystallites of monoclinic modification.  相似文献   

14.
We have developed a growth procedure for realizing a low defect density GaP layer on an Si substrate. The growth procedure consists of two parts. One is the post-growth annealing for the annihilation of stacking faults (SFs). We have investigated an annihilation mechanism with molecular beam epitaxy grown GaP layers. 1-monolayer-thick SFs typically generate from the GaP/Si interface in a non-annealed GaP layer. In a 700 °C annealed GaP layer, generation points of these SFs tend to shift toward the GaP surface. In a 730 °C annealed GaP layer, SFs density is effectively decreased. These results suggest that SFs are annihilated through the climb motion of two partial dislocations during the post-growth annealing. Another one is the optimized shutter sequence for migration enhanced epitaxy. We have revealed that it is effective for the suppression of both three-dimensional growth and melt-back etching to increase in a stepwise manner the number of supplied Ga atoms per cycle. As a result, the generation of threading dislocations and pits is remarkably suppressed. A root mean square surface roughness of 0.13 nm is obtained within the critical thickness. We have estimated etch pit density (EPD) to be ∼7×105 cm−2 with a GaPN/GaP/Si structure. To the best of our knowledge, this value is same as that of commercially available GaP substrates and is the lowest one in the EPD of GaP/Si heteroepitaxy.  相似文献   

15.
The vapour growth of InAs1-xPx layers has been carried out by the hydride process. The phosphorus rich part of the system (0.7 ? x ? 1) was especially investigated. Heteroepitaxial deposits of InAs1-xPx and InP have been performed on substrates such as InAs, GaAs and GaP. A systematic study of the influence of the substrate orientation on the quality of the layer has been carried out by growth on hemispherical substrates. Preferential planes have been pointed out: (100) and (111) A for InAs, (111) for GaAs and GaP. The band gap variation as a function of the composition has been determined by photoluminescence at 4.2 °K and X-ray diffraction measurements. It fits the equation: EG(x) eV = 0.425 + 0.722 x + 0.273 x2 at 4.2 °K.  相似文献   

16.
Our study of samples grown in different metalorganic chemical vapor deposition reactors and with different growth conditions reveals that V-pits are always present in (InxAl1?x)N films whatever the layer thickness and the InN content. V-pits are empty inverted pyramids terminating threading dislocations. InN-rich triangular regions are present around the threading dislocations terminated by pits with a hexagonal 6-fold symmetry distribution in {11?20} planes. The nature of the facets of the V-pits depends on the growth conditions: pits with either {11?2l}, l being between 1 and 3, or {1?101} facets have been observed. Moreover, the nature of the threading dislocations terminated by pits also depends on the growth conditions. Our observations suggest that with a high V/III ratio only edge a+c-type dislocations are terminated by pits whereas with a low V/III ratio both edge a-type and mixed a+c-type dislocations are terminated by pits.  相似文献   

17.
SrTiO3/BaZrO3 heterofilms as buffer layers are deposited on (0 0 1) MgO substrates by an RF-sputtering technique. The atomic structure and the defect configuration at the interfaces are investigated by means of aberration-corrected high-resolution transmission electron microscopy. At the BaZrO3/MgO interface, two types of interfacial structures, MgO/ZrO2-type and MgO/BaO-type, are observed. Antiphase boundaries and dislocations are found at the BaZrO3/MgO interface. The formation of these lattice defects is discussed in terms of film growth and structural imperfections of the substrate surface. At the SrTiO3/BaZrO3 interface, a high density of misfit dislocations is observed with different configurations. The formation of these dislocations contributes both to the relaxation of the large misfit strain and to stopping of the further propagation of lattice defects which are formed in the BaZrO3 layer into the SrTiO3 layer.  相似文献   

18.
MgxZn1?xO has been used in various photovoltaic cells because its energy bandgap can be tailored by controlling the Mg composition in this ternary compound. The MgxZn1?xO layers with different surface morphologies including two-dimensional (2-D) films and one-dimensional (1-D) nanostructures are preferred for conventional p–n junction solar cells and polymer–inorganic hybrid solar cells, respectively. The MgxZn1?xO layers are sequentially grown on Ga-doped ZnO (GZO) transparent conductive electrode using metalorganic chemical vapor deposition (MOCVD). The effect of the buffer layers on MgxZn1?xO surface morphology is investigated. It is observed that MgxZn1?xO deposited at ~500 °C on a low-temperature (~250 °C) ZnO buffer layer is in the form of 2-D dense and smooth films, whereas, on a high-temperature (~520 °C) ZnO buffer layer is in the form of 1-D nanostructures. Based on the structure characterization results, a growth mechanism in terms of nucleation and texturing is proposed to explain the buffer layer effect.  相似文献   

19.
《Journal of Non》2006,352(9-20):1126-1129
We fabricated a series of a-SiNx/nc-Si/a-SiNx double-barrier structures by plasma-enhanced chemical vapor deposition and subsequent thermal annealing technique. The photographs of transmission electron microscopy show that the nc-Si layer has been formed. The estimated density of nc-Si dots is of the order of 1011–1012 cm−2. In the CV measurements, for the samples with thicker SiNx layer (30 nm), we observed CV hysteresis characteristics, which can be explained by charging effect in nc-Si through the F–N tunneling mechanism; while for the sample with thinner SiNx layer (5 nm), the CV curves show the peak structures which can be attributed to the resonant tunneling of electrons into nc-Si controlled by Coulomb blockade effect. From the interval bias voltage between the two peaks, the Coulomb charging energy of nc-Si dot was estimated.  相似文献   

20.
Optical and structural properties of tensile strained graded GaxIn1−xP buffers grown on GaAs substrate have been studied by photoluminescence, X-ray diffraction, atomic force microscopy, and scanning electron microscopy measurements. The Ga composition in the graded buffer layers was varied from x=0.51 (lattice matched to GaAs) to x=0.66 (1% lattice mismatch to GaAs). The optimal growth temperature for the graded buffer layer was found to be about 80–100 °C lower than that for the lattice matched GaInP growth. The photoluminescence intensity and surface smoothness of the Ga0.66In0.34P layer grown on top of the graded buffer were strongly enhanced by temperature optimization. The relaxation of tensile GaInP was found to be highly anisotropic. A 1.5 μm thick graded buffer led to a 92% average relaxation and a room temperature photoluminescence peak wavelength of 596 nm.  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