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 23 毫秒
1.
Hu Hefang  J.D. Mackenzie   《Journal of Non》1986,80(1-3):495-502
The effect of oxide impurity on the physical properties of 62ZrF4---8LaF3---30BaF2 (mol.%) glass was studied by equimolecular substitution of BaO for BaF2. It is shown that the oxide impurity decreases the infrared transparency beyond 6 μm, shifts the transmission cut-off wavelength to higher frequencies and causes an additional absorption shoulder at 1350 cm−1. The oxide impurity also increases the glass transition temperature, the crystallization temperature and the viscosity of the melt. The additional infrared absorption of oxide impurity in the fluorozirconate glasses results from the multiphonon process of the vibration of F---Zr---O bonds at 680 cm−1.  相似文献   

2.
An atomic structure of Al55(Cr1−xMnx)15Si30 (x = 0, 0.49,1) metallic glasses was studied by neutron diffraction. An advantage of the neutron diffraction technique was fully exploited by utilizing the negative scattering length of Mn to form a neutron zero scattering ‘alloy’ for the component Cr0.51Mn0.49 in this quaternary Al---(Cr, Mn)---Si alloy. This allows the atomic distribution of the resulting quasibinary Al---Si metallic glass to be derived directly. Moreover, the (Al, Si)---TM (TM = Mn, Cr) and TM---TM pair correlations were also extracted by taking appropriate linear combinations of the atomic structures for the Al55(Cr1−xMnx)15Si30 (x = 0, 0.49, 1) metallic glasses. A sharp first peak in the (Al,Si) ---TM pair correlations thus obtained led to the conclusion that a strong attractive interaction exists between (Al, Si) and TM atoms and, hence, that the presence of the TM atoms is responsible for the formation of an amorphous phase.  相似文献   

3.
Glasses in the quasi-binary system (As4S6)x(P4S10)1−x x = 0.1, …, 1.0, are produced and studied by thermal analysis, X-ray, and Raman spectroscopy. The phase diagram of the system and the critical cooling rate for glass formation both have their maximum at x = 0.5, corresponding to the compound As2P2S8; the X-ray structure of recrystallized samples can be described as a sum of the As2P2S8 (x = 0.5)- and the P4S10-structure (As4S6 not visible); Raman spectra of the glasses are again sums of As2P2S8- and As4S6/P4S10-spectra. All these observations support the assumption that a stable building block corresponding to the 1:1 compound As2P2S8 and surplus As4S6 (or P410) are the essential elements of the structure in the glasses.  相似文献   

4.
An attempt is made to calculate the compositional dependence of the optical gap (Eg) in Ge1−xSx, Ge40−xSbxS60 and (As2S3)x(Sb2S3)1−x non-crystalline systems in an alloy-like approach. From the comparison of both the experimental dependence of Eg and the calculated ones using the Shimakawa relation [Eg AB = YEg A + (1−Y)Eg B] it is assumed that this equation is useful for such systems or parts of the systems which behave like almost ideal solutions.  相似文献   

5.
The elastic properties of GexAsySe100−xy (0x30; 10y40) glasses have been studied. The results were analyzed in terms of the dependence on the theoretical mean coordination number (mean number of covalent bonds per atom) m (m=2+(2x+y)×0.01). Three ranges of m (2.1m2.51, 2.51<m2.78, 2.78<m3) were revealed, where different dependencies of elastic moduli (Young’s modulus, shear modulus) and Poisson’s ratio of glasses on m were observed.  相似文献   

6.
The La L1 and L3 XANES and L3 EXAFS have been investigated for the series of glasses 10K2O---50SiO2---x La2O3 (x = 1, 5, 10) and (10 − x)K2O---40SiO2−(x/3)La2O3 (x = 7.5, 5, 2.5) and model compounds La2O3, LaAlO3, LaPO4, La2NiO4, La2CuO4 and La(OH)3. An edge resonance at 25 eV above the L1 edge in the glass spectra is concentration-dependent, decreasing in intensity with increasing lanthanum concentration. The 2s → nd forbidden transition increases with La2O3 concentration, indicating a reduction in the ‘average’ site symmetry of the first coordination shell of La. Mapping X(k) space, which is a new and promising technique, was employed to extract bond distance, coordination number and thermal parameters from the EXAFS. By this method, one calculates the complete X(k) space a function of all physically reasonable values of the adjusted parameters in all possible combinations. The advantage in this method is the assurance of a global minimum. Bond lengths were comparable to those obtained by Fourier transforming the phase corrected EXAFS. The values are 2.42 Å (± 0.03 Å) for La---O. The coordination numbers (N ≤ 7 ± 1.5) were derived by mapping and comparison to the published structures for other La compounds. X(k) mapping is compared with least-squares fitting the data, and the correlation between the Debye-Waller factor and coordination number is also discussed.  相似文献   

7.
Q. Ma  D. Raoux  S. Benazeth 《Journal of Non》1992,150(1-3):366-370
The structures of AsxTe1−x(x = 0.2, 0.4, and 0.5) glasses are studied using the differential X-ray anomalous scattering technique. The partial distribution functions have also been obtained for the stoichiometric As2Te3 glass, giving information on the medium range ordering. All the glasses show chemical disorder to differing extents, the As2Te3 glass being the most disordered. The Te coordination number undergoes a change at x=0.4 where it is 2.5 compared with 2 for AsTe. This change indicates the existence of about 50% of threefold Te sites in the stoichiometric glass, as well as in the Te-rich glasses. Some of the physical properties of the glasses may be explained based on these results.  相似文献   

8.
The electrical conductivities of (1−x) Li2O · x BaO · 2 SiO2, (1−x) Na2O · x MgO ·2 SiO2, (1−x) Na2O · x CaO · SiO2 and (1−x) Na2O · x BaO · 2SiO2 glasses were measured at temperature ranging from room temperature to 450°C. The transport numbers for Na+ ion in (1−x) Na2O · x BaO · 2 SiO2 glasses were measured. It was found that the alkali ion carried a significant part of the current in these glasses except one that had no alkali ions, and the conductivity decreased markedly as the alkali oxide was substituted by an alkaline earth oxide. The results of conductivity measurements combined with the data hitherto reported on mixed alkali glasses led to the proposal that the so-called mixed alkali effect could be explained on the basis of the independent path model in which it is assumed that cations can move only through vacant sites left by those of the same type.  相似文献   

9.
The 10 μm transparency of halides addresses directly its oxide content. The conversion from oxide to fluoride must have ΔG0. HF(g) is different from the other hydrogen halides because its ΔG of formation is larger than that of H2O(g). Decrease in the hydrolysis of RF3 (R=rare earth) influences the crystal phase stability and the tendency of the crystal to crack. The base-to-acid behavior of +3 rare earths provides quantitative support. HF(g) and RAP (reactive atmosphere process) CF4(g) formations of YF3 from Y2O3 are given in detail. This indifference to HF(g) is reported also in the congruent growth of CaF2, SrF2, and BaF2. The explanation is seen in ZrF4, HfF4, and ThF4. Although RAP does not use HF(g), it permits the congruent growth of good-quality single crystals with very low H2O(g) content.  相似文献   

10.
We studied the structural and optical properties of a set of nominally undoped epitaxial single layers of InxGa1−xN (0<x0.2) grown by MOCVD on top of GaN/Al2O3 substrates. A comparison of composition values obtained for thin (tens of nanometers) and thick (≈0.5 μm) layers by different analytical methods was performed. It is shown that the indium mole fraction determined by X-ray diffraction, measuring only one lattice parameter strongly depend on the assumptions made about strain, usually full relaxation or pseudomorphic growth. The results attained under such approximations are compared with the value of indium content derived from Rutherford backscattering spectrometry (RBS). It is shown that significant inaccuracies may arise when strain in InxGa1−xN/GaN heterostructures is not properly taken into account. Interpretation of these findings, together with the different criteria used to define the optical bandgap of InxGa1−xN layers, may explain the wide dispersion of bowing parameters found in the literature. Our results indicate a linear, Eg(x)=3.42−3.86x eV (x0.2), “anomalous” dependence of the optical bandgap at room temperature with In content for InxGa1−xN single layers.  相似文献   

11.
Heavily carbon-doped p-type InxGa1−xAs (0≤x<0.49) was successfully grown by gas-source molecular beam epitaxy using diiodomethane (CH2I2), triethylindium (TEIn), triethylgallium (TEGa) and AsH3. Hole concentrations as high as 2.1×1020 cm−3 were achieved in GaAs at an electrical activation efficiency of 100%. For InxGa1−xAs, both the hole and the atomic carbon concentrations gradually decreased as the InAs mole fraction, x, increased from 0.41 to 0.49. Hole concentrations of 5.1×1018 and 1.5×1019 cm−3 for x = 0.49 and x = 0.41, respectively, were obtained by a preliminary experiment. After post-growth annealing (500°C, 5 min under As4 pressure), the hole concentration increased to 6.2×1018 cm−3 for x = 0.49, probably due to the activation of hydrogen-passivated carbon accepters.  相似文献   

12.
AC conductivity and dielectric relaxation measurements of the bulk amorphous compositions in the pseudo-binary system (As2S3)1-x(PbS)x (x = 0, 0.1, 0.4 and 0.5) in the frequency range 500 Hz-10 kHz and in the temperature span 180–450 K are reported. The temperature dependence of the ac conductivity, σac(ω), has a broad structure at all frequencies in compositions with x = 0.1, 0.4 and 0.5 whose peak position is not thermally activated. A similar structure was also observed in the data on the dielectric constant, ε1, which peaked at a frequency of 1 kHz in the composition with x = 0.5. Analysis of the results using the correlated barrier hopping model revealed that the electronic conduction takes place by single polaron and bipolaron hopping processes at high and low temperatures, respectively, in compositions containing Pb. The microstructure and phase-separation in the glasses containing Pb influence the electrical transport and dielectric dispersion. This study has revealed the possible presence of a phase transformation at around 300 K at a frequency of 1 kHz in the dielectric dispersion behaviour of composition with x = 0.5.  相似文献   

13.
The 11B, 27Al, 29Si and 31P magic angle spinning (MAS) NMR spectra of MO–P2O5, MO–SiO2–P2O5 and MO(M2O)–SiO2–Al2O3–B2O3 (M=Mg, Ca, Sr and Ba, M=Na) glasses were examined. In binary MO–P2O5 (M=Ca and Mg) glasses, the distributions of the phosphate sites, P(Qn), can be expressed by a theoretical prediction that P2O5 reacts quantitatively with MO. In the ternary 0.30MO–0.05SiO2–0.65P2O5 glasses, the 6-coordinated silicon sites were detected, whose population increases in the order of MgOxCaO–0.05SiO2–(0.95−x)P2O5 glasses, its population increases with an increase in f (=([P2O5]−[MO]−[B2O3]−[Na2O])/[SiO2]) and has maximum at f=9. The signal due to the 5-coordinated silicon atoms is also observed when x is smaller than 0.45. When three network-forming oxides such as SiO2, Al2O3 and B2O3 coexist, Al2O3 reacts preferably with MO. The populations of 4-coordinated boron atoms, N4, are expressed well with r/(1−r), where r=([Na2O]−[Al2O3])/([Na2O]−[Al2O3]+[B2O3]). The correlation of the Raman signal at 1210 and 1350 cm−1 with the NMR signal of Si(Q6) at −215 ppm is also seen.  相似文献   

14.
The layers of ZnSe1−xTex (0 < x < 1.0) solid solutions have been grown by liquid-phase epitaxy in a closed tube at 620–680 °C. Zinc chloride served as a solvent. ZnTe and ZnSe crystals were used as sources and substrates with orienting surfaces (110) and (111) for ZnSe and (110) for ZnTe. The composition of the grown layer was specified by the relative content of the ZnSe and the ZnTe in the solution and was controlled by X-ray analysis. The position of the exciton bands in the photoluminescence spectra of ZnSe1−xTex over the interval 0.3 < x < 1.0 is in agreement with the free exciton energies calculated for these compositions. Relatively low-ohmic (of about 102 Ω cm) epitaxial layers of ZnSe1−xTex solid solutions were grown.  相似文献   

15.
R. Mathai  G. H. Frischat   《Journal of Non》1999,260(3):175-179
A glass of composition 53ZrF4–20BaF2–4LaF3–3AlF3–20NaF (Tg=260°C) was prepared by careful crucible melting. High-resolution atomic force microscopy of fracture surfaces displayed the presence of nano-pores with diameters of 20–50 nm, being 4–10 nm deep, in all glasses. It was further found that only glasses without annealing and glasses with an annealing step considerably below Tg showed a distinct pattern, i.e. ripples of ≈20 nm in diameter and an rms roughness of ≈0.6 nm. Glasses annealed either near Tg or at the temperatures of maximum nucleation or maximum crystal growth rates showed both regions with the ripple pattern and regions with nano-hillocks, growing in size with increasing annealing temperature and time. Thus these hillocks nearly reach micro-dimensions of ≈270 nm in diameter and ≈65 nm in height following a 90 min annealing step at 343°C, the temperature of maximum crystal growth. These findings give evidence that the glass system, which is thought to be one of the most suitable for fiber drawing, is much less stable against nucleation and crystallization than anticipated.  相似文献   

16.
The internal friction of xNa2O·(0.5−x)V2O5·O.5P2O5(x = 0.025–0.3) glasses was studied using the low-frequency torsion pendulum technique. The temperature spectrum of internal friction reveals three maxima. Maximum 1, the so-called “electron” maximum, is the same as observed in binary vanadium-phosphate glasses. The origin of maximum 2 can be attributed to ion migration. Maximum 3 appears for glasses containing more than 10 mol.% Na2O and is probably connected with sodium-proton interactions.  相似文献   

17.
Carbon-doped InxGa1−xAs layers (x=0−0.96) were grown by metalorganic molecular beam epitaxy (MOMBE) using trimethylgallium (TMG), solid arsenic (As4) and solid indium (In) as sources of Ga, As and In, respectively. The carrier concentration is strongly affected by growth temperature and indium beam flux. Heavy p-type doping is obtained for smaller In compositions. The hole concentration decreases with the indium composition from 0 to 0.8, and then the conductivity type changes from p to n at x=0.8. Hole concentrations of 1.0×1019 and 1.2×1018 cm-3 are obtained for x=0.3 and 0.54, respectively. These values are significantly higher than those reported on carbon-doped InxGa1−xAs by MBE. Preliminary results on carbon-doped GaAs/InxGa1−xAs strained layer superlattices are also discussed.  相似文献   

18.
J.W Park  Haydn Chen 《Journal of Non》1980,40(1-3):515-525
The infrared absorption spectra of sodium-disilicate glasses containing various amounts of Fe2O3 ([Na2O · 2 SiO2]1−x [Fe2O3]x, where X = 0.05, 0.1 and 0.2) were investigated in the wavenumber range from 200–2000 cm−1. The addition of Fe2O3 to the sodium-disilicate glass does not seem to introduce any new absorption band as compared with the spectrum of a pure sodium-disilicate glass; nevertheless, a general shift of the existing absorption bands toward lower wavenumbers is observed. The amount of shift is, in fact, proportional to the content of Fe2O3 in the glass. This observation is consistent with the recently proposed structural model for the bonding of Fe3+ ions in the iron-sodium-silicate glass system.

Annealing of 20 mol% iron oxide glasses at 550 and 580°C produced an extra sharp infrared absorption peak at about 610 cm−1 wavenumber. This new peak is believed to be related to the crystallized particles of the glass as concluded from both a scanning electron micrograph and an electron diffraction pattern.  相似文献   


19.
A series of lanthanide oxides was incorporated in a vitreous phosphate host network. Molar constituents of the glasses were typically (La2O3)10(RxOy)10(Al2O3)5(P2O5)75. Each glass had a different lanthanide (R atom) from the series; La, Ce, Pr, Nd, Sm, Eu, Tb, Dy, Ho, Er and the values of x and y depended on the valency of the rare-earth atom. Both X-ray and neutron diffraction were employed in examining their structures. The results indicate that the basic PO4 tetrahedral unit remains unaltered with an average P–O distance of and predominant Q2 linkages to its neighbouring units so as to form a continuous network while accommodating the included lanthanides. In accordance with this model, the average distance of rare-earth (comprising La and a second type of R atom) to oxygen decreased from 2.44 to 2.26 , a trend to be expected from the lanthanide contraction. The average oxygen coordination around the rare-earth was found to vary in the range of 6–8. With these average parameters, a small (74 atom) hand-built model was made to check the feasibility of constructing a continuous random network. Optical transmission measurements show all these glasses to absorb strongly in the UV region and to have marked absorption resonances in the visible region of 400–1000 nm except for the La, Ce, Eu, Tb containing glasses which have low or negligible absorption in the latter range.  相似文献   

20.
Raman spectra have been measured for ZnCl2---ZnX2 and ZnCl2---KX (X = Br, I) glasses to investigate the structure of the glasses with varying composition. The assignment of each band was made, and the change of the spectra with composition was explained in terms of the bridging and non-bridging states of halide ions and the change of the tetrahedral units, ZnXnCl4−n2− (n = 0–4), formed in the glasses. As the content of ZnX2 in ZnCl2---ZnX2 glasses increases (20 → 80 mol%), the peak frequency of the Zn---Cl stretching mode increases (238 → 248 cm−1 in X = I glasses, 238 → 259 cm−1 in X = Br glasses) while the Zn---I and Zn---Br stretching frequencies decrease (173 → 120 cm−1 for Zn---I, 196 → 157 cm−1 for Zn---Br). The decrease of the Zn---I and Zn---Br band frequencies was attributed to the increase of the number n of the ZnXnCl4−n2− tetrahedra. The increase of the Zn---Cl frequency suggests the existence of the bonding state of Cl ions which is intermediate between the bridging and the non-bridging states. In ZnCl2---KX glasses, the Zn---Clnon-bridging band at about 300 cm−1 was observed in addition to the bands observed in ZnCl2---ZnX2 glasses. The addition of KX produces non-bridging anions while the tetrahedral units, ZnXnCl4−n2− are also formed.  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