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 62 毫秒
1.
We report on the structural details and microphase separation of the bulk glasses Agx·(As33S67)100-x for 0x25. Glass–glass phase separation occurs over a wide range of Ag content, i.e. 4x20. An off-resonant polarized Raman spectroscopic study has been carried out to elucidate structural aspects at the short- and medium-range structural order of the glasses. Analysis of Raman spectra revealed quantitative changes of the sulfur-rich microenvironments that reduce upon adding Ag. Scanning electron microscopy combined with X-rays microanalysis have been utilized to examine the type and extent of phase separation, and to provide quantitative details on the atomic concentrations in the Ag-poor and Ag-rich phases. It has been shown that at 7 at.% Ag the Ag-rich phase percolates through the structure; this effect can be associated with an ionic-to-superionic behavior of these glasses in accordance with similar studies on the stoichiometric arsenic sulfide glass; although the phase separation observed in the present glasses is qualitatively different.  相似文献   

2.
The role of the composition and of the related changes of the structure in the formation of the surface of amorphous AsxSe1−x (0 < x < 0.5) layers before and after light treatment was investigated by direct measurements of the surface roughness at nanometer-scale and surface deformations at micrometer-scale under influence of illumination. It was established that the surface roughness of the films, deposited by vacuum thermal evaporation, decreased with increasing As content, especially in compositions 0.1  x  0.3, where the maximum light stimulated surface deformations (localized expansion) occurs. Both relate to the rigidity percolation range and the maximum photoplastic effects, which are not directly connected to the known photodarkening effect, since it is minimal for these compositions.  相似文献   

3.
Growth on AlN/4H–SiC substrates of coalesced, non-polar GaN films having volumes of material with reduced densities of dislocations and stacking faults has been achieved from etched stripes via the statistical and experimental determination of the effect of temperature and V/III ratio on the lateral and vertical growth rates of the GaN{0 0 0 1} faces combined with pendeo-epitaxy. AFM of the uncoalesced GaN(0 0 0 1) and GaN vertical faces revealed growth steps with some steps terminating at dislocations on the former and a pitted surface without growth steps, indicative of decomposition, on the latter. Coalescence was achieved via (a) a two-step route and the parameters of (1) and V/III=1323 for 40 min and (2) 1020 °C and V/III=660 for 40 min and (b) a one-step route that employed and a V/III ratio=660 for 6 h. The densities of dislocations in the GaN grown vertically over and laterally from the stripes were 4×1010 cm−2 and 2×108 cm−2, respectively; the densities of stacking fault in these volumes were 1×106 cm−1 and 2×104 cm−1, respectively. The defects in the wing material were observed primarily at the bottom of the film where lateral growth of the GaN occurred from the AlN and the SiC. Plan view AFM also revealed different microstructures and a reduction in the RMS roughness values from 1.2 to 0.95 nm in these respective regions.  相似文献   

4.
Cuprous oxide (Cu2O) thin films were grown epitaxially on c-axis-oriented polycrystalline zinc oxide (ZnO) thin films by low-pressure metal organic chemical vapor deposition (MOCVD) from Copper(II) hexafluoroacetylacetonate [Cu(C5HF6O2)2] at various substrate temperatures, between 250 and 400 °C, and pressures, between 0.6 and 2.1 Torr. Polycrystalline thin films of Cu2O grow as single phase with [1 1 0] axis aligned perpendicular to the ZnO surface and with in-plane rotational alignment due to (2 2 0)Cu2O(0 0 0 2)ZnO; [0 0 1]Cu2O[1 2¯ 1 0]ZnO epitaxy. The resulting interface is rectifying and may be suitable for oxide-based p–n junction solar cells or diodes.  相似文献   

5.
Amorphous non-hydrogenated germanium carbide (a-Ge1 − xCx) films have been prepared by magnetron co-sputtering system designed by ourselves. The chemical bonding and microstructure have been analyzed using X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectroscopy (FTIR) and Raman spectroscopy. The optical properties of the films have been investigated by means of spectroscopic ellipsometry. The relationship between the chemical bonding and the optical properties has been explored. It has been found that all films with the constant carbon content are amorphous. The sp2 CC and sp3 GeC bonds increase with Ts, and some sp2 CC bonds gain infrared activity. The fraction of sp3 GeC bonds rises with Ts, but the fraction of sp3 GeGe bonds gradually drops down. In addition, the refractive index and extinction coefficient increase with Ts. The film optical gap is seen to reach 1.15 eV when Ts is 200 °C. However, the optical properties of a-Ge1 − xCx films almost remain stable with the substrate temperature.  相似文献   

6.
A.N. Trukhin  K.M. Golant 《Journal of Non》2009,355(34-36):1719-1725
Photoluminescence (PL) spectra and kinetics of high purity amorphous silicon dioxide with ultra low hydroxyl content is studied under the excitation by F2 excimer laser (157 nm wavelength) pulses. Materials synthesized in the SPCVD plasma chemical process are studied before and after fusion. Two bands are found in the PL spectra: one centered at 2.6–2.9 eV (a blue band) and the other at 4.4 eV (a UV band). Luminescence intensity of unfused material is found to increase significantly with exposure time starting from a very small level, whereas in fused counterpart it does not depend on irradiation time. Both bands show complicated decay kinetics, to which add exponential and hyperbolic functions. The UV band of the unfused material is characterized by decay with exponential time constant τ  4.5 ns and hyperbolic function tn, where n = 1.5 ± 0.4. For the blue band the hyperbolic decay kinetics with n  1.5 extends to several milliseconds, gradually transforming to the exponential one with τ = 11 ± 0.5 ms. In fused glass relative contribution of the fast component to the UV band is small whereas for the blue one it is great, that allows one to more accurately determine the hyperbolic law factor n = 1.1 ± 0.1 typical for tunneling recombination. Simultaneous intracenter and recombination luminescence, the later occurring with the participation of laser radiation induced defects, add particular features to the decay kinetics. Spectra of the above luminescence processes are different. A less sharp position of bands is associated with the recombination luminescence. The origin of the observed PL features we attribute to the presence of oxygen deficient centers in glass network in the form of twofold coordinated silicon. Such centers being affected by network irregularities can be responsible for the recombination PL component. A great variety of network irregularities is responsible for centers’ structural inequivalence, which causes a non-uniform broadening of PL spectral and kinetic parameters.  相似文献   

7.
B. Frumarova  M. Frumar  J. Oswald  M. Kincl  M. Vlcek 《Journal of Non》2009,355(37-42):1865-1868
Glasses of systems 100-y((GeS2)80(Sb2S3)20−x(PbI2)x)yPr2S3, x = 0; 2; 5, 8; y = 0; 0.01; 0.1; 0.5 and 99.9-z((GeS2)80(Sb2S3)18(PbI2)2)0.1Pr2S3zYb2S3, z = 0.05; 0.1; 0.15) were synthesized in high purity. Optically well transparent glasses were obtained for x  5 mol.% PbI2, for y  0.1 mol.% Pr2S3 and for z  0.15 mol.% Yb2S3. The glasses were stable and homogeneous, as confirmed by X-ray diffraction and electron microscopy, with high optical transmittivity from visible (red) region up to infrared region (900 cm−1). The density of the glasses was 3.26–3.33 gcm−3 for PbI2 containing glasses. The glass transition temperature, Tg, was 320–336 °C. The optical absorption bands in rare-earth doped glasses corresponded to 3H43F4, 3H43F3, 3H4–(3F2 + 3H6) f–f electron transitions of Pr3+ ions and to 2F7/22F5/2 f–f electron transitions of Yb3+ ions. Strong luminescence band with maximum near 1340 nm (electron transition 1G43H5) was found in Pr2S3 doped glasses. The intensity of this band was rising with doping by Yb3+ ions. The possible mechanism of the luminescence enhancement is suggested.  相似文献   

8.
The growth of highly oriented 3C–SiC directly on an oxide release layer, composed of a 20-nm-thick poly-Si seed layer and a 550-nm-thick thermally deposited oxide on a (1 1 1)Si substrate, was investigated as an alternative to using silicon-on-insulator (SOI) substrates for freestanding SiC films for MEMS applications. The resulting SiC film was characterized by X-ray diffraction (XRD) with the X-ray rocking curve of the (1 1 1) diffraction peak displaying a FWHM of 0.115° (414″), which was better than that for 3C–SiC films grown directly on (1 1 1)Si during the same deposition process. However, the XRD peak amplitude for the 3C–SiC film on the poly-Si seed layer was much less than for the (1 1 1)Si control substrate, due to slight in-plane misorientations in the film. Surprisingly, the film was solely composed of (1 1 1) 3C–SiC grains and possessed no 3C–SiC grains oriented along the 3 1 1 and 1 1 0 directions which were the original directions of the poly-Si seed layer. With this new process, MEMS structures such as cantilevers and membranes can be easily released leaving behind high-quality 3C–SiC structures.  相似文献   

9.
Chalcogenide bulk glasses Ge20Se80−xTex for x(0,15) have been prepared by systematic replacement of Se by Te. Selected glasses have been doped with Ho, Er and Pr, and samples have been characterized by transmission spectroscopy, measurements of dc electrical conductivity and low-temperature photoluminescence. Absorption coefficients have been derived from measured transmittance and estimated reflectance. Arrhenius plots of dc electrical conductivity, in the measured temperature range 300–460 K, are characterized by single activation energies roughly equal to the half of the optical gap. Activation energies deduced from Arrhenius plots reveal a systematic decrease with increasing Te content. Similarly, both absorption and low-temperature photoluminescence spectra reveal shifts of absorption edge and/or dominant luminescence band to longer wavelength due to Te → Se substitution. Samples doped with Ho and Er exhibit a strong luminescence at 1200 and 1540 nm due to 5I6 → 5I8 and 4I13/2 → 4I15/2 transitions of Ho3+ and Er3+ ions, respectively. Pr doped samples exhibit only a relatively weak luminescence peak at 1590 nm, which we tentatively assign to 3F3 → 3H4 transition of Pr3+ ions. Absorption of the base glass luminescence at 1460 and 1520 nm has been observed at low temperature on samples doped with Pr and Er, respectively.  相似文献   

10.
A previews study of germanium selenide glass films by scanning electron microscopy and atomic force microscopy revealed a heterogeneous surface morphology consisting of granular regions 15–50 nm in size, which cause high optical losses. The present work was performed in order to further characterize such materials using spectroscopic ellipsometry, infrared (IR) and Raman spectroscopies. Chalcogenide glass films with GeSe2, Ge28Sb12Se60 and GeSe compositions have been deposited on single crystal silicon and silica glass substrates by vacuum thermal evaporation. The film thickness and the optical constants were obtained from spectroscopic ellipsometry using the Tauc-Lorenz dispersion formula. A model was derived for the film structure, which included a roughness layer at the surface. This top layer was found to have a thickness of 5–15 nm, of the order of the size of the granular regions previously reported. The optical bandgap of the samples increased with increasing selenium content, while the refractive index decreased. Despite a previous report of large scale phase separation in bulk Ge26Sb14Se60 glass, the fundamental IR and Raman spectra obtained in the present work did not provide any clear evidence for such phase separation which could be associated with the heterogeneous nanostructure observed at the surface of the films.  相似文献   

11.
Nd3+-doped NaGd(MoO4)2 crystal with dimensions were grown by Czochralski method. Nd3+:NaGd(MoO4)2 crystal melts at 1182 °C. The hardness of Nd3+:NaGd(MoO4)2 crystal is 334 VDH. The specific heat is 72.6 cal/mol K. The thermal expansion coefficients are for c-axis and for a-axis, respectively. The absorption cross-sections of Nd3+:NaGd(MoO4)2 crystal are with a FWHM of 9 nm at the 804 nm for π-polarization and with a FWHM of 17 nm at 807 nm for σ-polarization, respectively. The emission cross-section σem are at 1063 nm for π-polarization and 1.94×10-20 at 1070 nm cm2 for σ-polarization, respectively. The fluorescence lifetime τf is 93.9 μs at room temperature.  相似文献   

12.
The chalcogenide multilayers were prepared as dielectric mirrors having the first order stop bands in the near infrared region 1.55 μm. The 7.5 layer pairs of the alternating amorphous Sb–Se and As–S layers were deposited on glass substrates using a conventional thermal evaporation method. To center the stop bands of the 15-layer dielectric mirrors at 1.55 μm, the layer thicknesses 117 nm for Sb–Se and 169 nm for As–S single layers were calculated from the quarter wave stack condition. The optical reflection and transmission spectra of the prepared mirrors were measured using a UV/VIS/NIR and FT-IR spectroscopy at the ambient and elevated temperatures. The optical reflection of the annealed 15-layer chalcogenide mirror was found higher than 99% in the range of 1440–1600 nm. As the 200 nm thick gold layer was added between the substrate and the chalcogenide mirror, the stop band of the annealed Au/multilayer system broadened to 1360–1740 nm simultaneously with an appearance of the 15% transmission peak at 1.55 μm. A preparation of similar metal/multilayer systems is one of the possible ways how to design the dielectric filters for near infrared region exploiting the good optical quality of the chalcogenide films and their simple deposition.  相似文献   

13.
The space group of tungstogallate acid, H5GaW12O40, reported by Niu et al. (J. Chem. Crystallogr. 2003, 33, 799) should be I–43m instead of Cm; the monoclinic C-centered cell is transformed to the I-centered cubic cell by (1 0 0; ,– , 1; – , – , –1).  相似文献   

14.
The tris-2-chloro and 2-bromotribenzylamines are prepared from aqueous ammonia and 2-chlorobenzyl chloride and 2-bromobenzyl bromide, respectively, in ethanol. Recrystallization yielded colorless cubes of each product. The crystal structures are each solved in space group P , and are isostructural. The tris-2-chloro compound, 1, has a = 7.4226(5) Å, b = 9.0825(7) Å, c = 14.529(1) Å, = 78.279(1), = 82.389(1), = 84.661(1), and V = 948.41(12) Å3 with Z = 2, and dcalc = 1.368 Mg/m3. The tris-2-bromo analog, 2, has a = 7.6569(11) Å, b = 9.0922(13) Å, c = 14.614(2) Å, = 79.286(2), = 81.777(2), = 85.401(2), and V = 987.9(2) Å3 with Z = 2, and dcalc = 1.762 Mg/m3. Lithium–halogen exchange experiments conducted in tetrahydrofuran at –78C using n-butyl lithium revealed that no exchange occurred for the tris-2-chloro compound, but did occur for the tris-2-bromo analog to yield tribenzylamine upon quench and work-up.  相似文献   

15.
16.
A single crystal of phase 1 of 1,2-difluoroethane was grown from the melt directly on an X-ray diffractometer close to the melting point of 169 K. It crystallizes in the monoclinic space group C2/c with lattice parameters a = 7.775(4), b = 4.4973(7), c = 9.024(3) Å, = 101.73(1)°, V = 308.9(2) Å3, d calc = 1.420 g cm–3 for Z = 4. A second phase of 1,2-difluoroethane was obtained under similar conditions which crystallizes in the orthorhombic space group P212121 with the unit cell parameters a = 8.0467(16), b = 4.5086(9), c = 8.279(2) Å,V = 300.36(11) Å3, d calc = 1.461 g cm–3 for Z = 4. In both phases the 1,2-difluoroethane molecules adopt the gauche conformation with F–C–C–F torsion angles close to 68°. Crystals of 1,2-diiodoethane C2H4I2 were grown from pentane at –30°C. A platelet single crystal of the size 0.35 × 0.25 × 0.03 mm was measured with Mo K-radiation at 153 K. 1,2-Diiodoethane crystallizes in the monoclinic space group P21/n with a unit cell of a = 4.6051(7), b = 12.939(2), c = 4.7318(7) Å, = 104.636(3)°, V = 272.79(7) Å3, Z = 2, d calc = 3.431 g cm–3, (MoK) = 11.353 mm–1. In the molecule the two neighboring iodine atoms are positioned anti. The shortest intermolecular contacts occur via iodine–iodine interactions resulting in layers of molecules in the crystal.  相似文献   

17.
We report a structural investigation of bulk Ge-rich Ge–S–AgI chalcohalide glasses. A vibrational spectroscopic study of the quaternary system (AgI)x (GeS1.5)100−x (0  xAgI  20) has been undertaken using infrared spectroscopy and Fourier transform Raman scattering. It was found that the GeS1.5 Raman spectrum is compatible with a glass structure composed of corner- and edge-sharing mixed GeSnGe4−n (n = 0–4) tetrahedra where units with n = 2–4 dominate, whilst the fraction of corner-sharing units are significantly lower than the corresponding fraction in the stoichiometric GeS2 glass. The addition of AgI has revealed a subtle but systematic effect in the structure of the Ge-rich glass matrix, manifested by mild decrease of the ES units and the concomitant increase of complex GeSnI4−n or GeSnGemI4-nm tetrahedra whose vibrational modes form a continuum at low frequencies. Although, AgI seems to cause subtle structural changes due to the formation of Ge–I bonds, it is also evident that AgI does not act as a real modifier that would depolymerize appreciably the Ge–S network structure.  相似文献   

18.
A new Er(III)–Na(I) coordination polymer of stoichiometry [NaEr2L5(H2O)6(NO3)](NO3)·3.5H2O (HL = picolinic acid N-oxide) has been synthesized and characterized by single-crystal X-ray analysis. Crystals are triclinic, P with a = 9.823(2), b = 12.453(2), c = 20.643(4) Å; = 98.49(3), ( = 101.40(3), = 108.69(3)°; V = 2284(1) Å3; Z = 2. Of the two independent eight-coordinate erbium(III) ions in this complex, one is surrounded by four bidentate chelating L ligands, and the other by one bidentate chelating L ligand, four aqua ligands and two anti-carboxylate oxygen atoms from two neighboring [ErL4] units. The sodium(I) ion is in a distorted octahedral environment, being coordinated by a unidentate nitrate anion, three aqua ligands and two anti-carboxylate oxygen atoms from two adjacent [ErL4] units. The complex is built from zigzag chains of syn-anti carboxylate-bridged erbium(III) moieties directed in the a direction, which are cross-linked pairwise by aqua-bridgeddimericsodium(I) units. The resulting composite polymeric chains are further connected by hydrogen bonds to form a three-dimensional network.  相似文献   

19.
Thermal and Me3NO-assisted activation of the donor–acceptor complex Ru2(CO)6(bpcd) (1) [where bpcd = 4,5-bis(diphenylphosphino)-4-cyclopenten-1,3-dione] with PMe3 or tBuNC affords the mono-substituted complexes Ru2(CO)5L(bpcd), as a result of regiospecific ligand attack at the diphosphine-substituted ruthenium center. Solution NMR measurements (1H and 31P) reveal that the PMe3 derivative exists as a noninterconverting mixture of axial (3a) and equatorial (3e) isomers, with the only the equatorial isomer being observed for Ru2(CO)5(tBuNC)(bpcd) (5). Near-UV irradiation of 1 in the presence of added ligand yields Ru2(CO)5L(bpcd), in addition to the known 2-phosphido complex Ru2(CO)6 [-C=C(PPh2)C(O)CH2C(O)](2-PPh2) (2) and the corresponding phosphido-substituted complexes Ru2(CO)5L[-{C =C(PPh2)C(O)CH2C}(O)]2-PPh2)[4 (L = PMe3); 6 (L = tBuNC)]. As with compounds 3a, 3e, and 5, both 4 and 6 exhibit ligand attachment at the diphosphine-substituted ruthenium center. The molecular structures of 3e, 4, 5, and 6 were determined by X-ray crystallography. 3e, as the 1/2 C6H6 solvate, crystallizes in the monoclinic space group C2/c: a = 40.573(3) Å, b = 10.2663(9) Å, c = 18.347(1) Å, = 95.371(6)°, V = 7609(1) Å3 and Z = 8; 4, crystallizes in the monoclinic space group P21/n: a = 10.8241(8) Å, b = 18.074(1) Å, c = 19.194(1) Å, = 96.968(6)°, V = 3727.3(5) Å3, and Z = 4; 5, as the 1/2CH2Cl2 solvate, crystallizes in the monoclinic space group C2/c: a = 40.955(3) Å, b = 9.7230(6) Å, c = 20.542(1) Å, = 106.596(5)°, V = 7839.2(9) Å3, and Z = 8; 6, as the 1/2C5H12 solvate, crystallizes in the monoclinic space group P21/c: a = 21.773(2) Å, b = 10.907(3) Å, c = 18.744(4) Å, = 114.68(1)°, V = 4045(1) Å3, and Z = 4. The site occupied by the PMe3 and tBuNC ligands in these compounds is discussed relative to the steric size/electronic properties of the ancillary ligand and its interaction with the bpcd ligand.  相似文献   

20.
Ten-vertex [6,6-(PMe2Ph)2-arachno-6-PdB9H12-9-(PMe2Ph)] 1a and eleven-vertex [7,7-(PMe2Ph)2-nido-7-PdB10H12] 2a have been isolated as occasional by-products from an extension to palladaborane chemistry of the one-pot route for the preparation of the nine-vertex platinaborane [(PMe2Ph)2PtB8H12] 3b from [PtCl2(PMe2Ph)2] and B10H14, but using [PdCl2(PMe2Ph)2] instead of [PtCl2(PMe2Ph)2] to generate [4,4-(PMe2Ph)2-arachno-4-PdB8H12] 3a in yields of upto 75%. The two by-products 1a and 2a are each characterized by single-crystal X-ray diffraction analysis. Space group and cell parameters are as follows: for 1a, triclinic, , a = 10.1139(2) Å, b = 10.3955(2) Å, c = 31.5086(7) Å, = 94.4720(12)°, = 91.6420(10)°, and = 105.3790(11)°; for 2a, monoclinic, P21/n, a = 14.6059(3) Å, b = 10.9988(2) Å, c = 15.1516(3) Å, and = 93.0000(14)°. Whereas 1a and 1b are isomorphous, Compounds 2a and 2b are not, and show significant differences in intramolecular conformation.  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