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 95 毫秒
1.
Epoxy coatings containing TiO2 nanocontainers were applied on aluminium alloy (AA) 2024-T3 for corrosion protection. The nanocontainers were loaded with the corrosion inhibitor 8-hydroxyquinoline (8-HQ). Epoxy coatings were deposited via the dip-coating process. The morphology of the coatings was examined by scanning electron microscopy (SEM). The composition of the films was determined by energy dispersive X-ray analysis (EDX). Electrochemical impedance spectroscopy (EIS) was employed for the characterization of the corrosion resistance of these coatings. The total impedance values were measured as a function of time exposure in corrosive environment. We observed a continuous increase of the total impedance value with the time of exposure suggesting a possible self-healing effect due to the release of the inhibitors from the nanocontainers. Furthermore, addition of loaded nanocontainers into the coatings leads to the enhancement of the barrier properties of the coatings. Conclusively, we observed an improvement of the performance of the coatings due to the loaded nanocontainers.  相似文献   

2.
The corrosion inhibition of AA2024-T4 in 3.5% NaCl solution by 8-hydroxylquinoline (8-HQ) was investigated by potentiodynamic polarisation (PDP), electrochemical impedance spectroscopy and dynamic electrochemical impedance spectroscopy. Experimental results were supported with scanning electron microscopy (SEM), atomic force microscopy and Fourier transform-infrared (FTIR) spectroscopy analysis. It was found that 8-HQ molecules adsorbed on the alloy surface and protected it against corrosion. SEM, energy dispersive spectroscopy, and FTIR results confirm the adsorption of 8-HQ molecules on AA2024-T4. The inhibition efficiency of 8-HQ is found to increase with increase in concentration and the highest concentration studied (0.05 M) offered corrosion inhibition efficiency of 84%. PDP results show that 8-HQ acts as mixed type inhibitor in the studied medium.  相似文献   

3.

The present study focuses on the synthesis of novel lanthanum cerium molybdate (LCM) nanoparticles by sol–gel synthesis method and their use in the development of nanocontainers in an anticorrosive coating application. The obtained nanoparticles were used as core material with two different polyelectrolytic shells comprising of polypyrrole (PPY) and polyacrylic acid (PAA) or polyethyleneimine (PEI) and polystyrene sulfonate (PSS) involving the entrapment of benzotriazole (BTA) as the corrosion inhibitor using layer-by-layer (LBL) deposition method. At each step of this nanocontainer synthesis, the thickness of the layers, surface charges and the presence of the functional groups were determined by particle size, zeta potential and Fourier transform infrared spectroscopy (FTIR) analysis, respectively. The X-ray diffractograms (XRD) indicated the change in the crystallinity of the nanoparticles and nanocontainers while thermogravimetric analysis (TGA) showed the thermal degradation behavior of the nanocontainers. The morphological studies conducted using scanning electron microscopy (SEM) exhibited the formation of nanocontainers containing nanoparticles in their cores. The release of BTA from the nanocontainers was evaluated at different pH values. The anticorrosive performance of the nanocontainers was examined by incorporation of the nanoparticles and nanocontainers in a commercial epoxy coating system and to be applied on mild steel and magnesium panels by electrochemical corrosion analysis. Tafel plots demonstrated the decrease in the current density with an increase in the loading percentage of nanocontainers in the epoxy system while Bode plots confirmed the significant improvement in the corrosion protection of the mild steel and magnesium by LCM nanoparticles and nanocontainers.

  相似文献   

4.
The sintering of spherical borosilicate glass powder (particle size 5 to 10 μm) under a uniaxial stress was studied at 800°C. The experiments allowed the measurement of the kinetics of densification and creep, the viscosities for creep and bulk deformation, and the sintering stress which was found to increase with density. The data show excellent qualitative agreement with Scherer's theory of viscous sintering. In addition, the quantitative comparison between theory and experiment shows good agreement; the measured viscosity of the bulk glass was ∽1×109 P (∽1×108 Pa·s) compared to ∽3×109 P (∽3 Pa·s) obtained by fitting the data with Scherer's theory.  相似文献   

5.
Hollow ceria nanospheres were synthesized using anionic polystyrene lattices which were prepared by emulsion polymerization of styrene using potassium persulfate as the initiator. These anionic colloidal particles were dispersed in water in the presence of poly(vinylpyrrolidone) and mixed with aqueous solutions of cerium (III) acetylacetonate [Ce(acac)3]. Subsequently, hollow nanospheres of cerium compounds were obtained by calcination of the coated polystyrene lattices at an elevated temperature in air. The hollow ceria nanospheres were characterized by scanning electron microscopy, transmission electron microscopy, Fourier transform infrared spectroscopy, X-ray diffraction, thermogravimetric analysis, and differential thermal analysis. The hollow ceria nanospheres were coated with conductive polymers (polyaniline and polypyrrole) via an electropolymerization process. Moreover, the antibacterial action of illuminated hollow ceria nanospheres and hollow ceria nanospheres coated with conductive polymers (CPCeO2) on a pure culture of Escherichia coli was studied. A decrease of E. coli concentration was observed after illumination of bacteria in the presence of hollow ceria nanospheres and CPCeO2.  相似文献   

6.
The thermal expansion of the hexagonal (6H) polytype of α-SiC was measured from 20° to 1000°C by the X-ray diffraction technique. The principal axial coefficients of thermal expansion were determined and can be expressed for that temperature range by second-order polynomials: α11= 3.27 × 10–6+ 3.25 × 10–9T – 1.36 × 10–12 T 2 (1/°C), and ş33= 3.18 × 10–6+ 2.48 × 10–9 T – 8.51 × 10–13 T 2 (1/°C). The σ11 is larger than α33 over the entire temperature range while the thermal expansion anisotropy, the δş value, increases continuously with increasing temperature from about 0.1 × 10–6/°C at room temperature to 0.4 × 10–6/°C at 1000°C. The thermal expansion and thermal expansion anisotropy are compared with previously published results for the (6H) polytype and are discussed relative to the structure.  相似文献   

7.
The deviation from stoichiometry, δ, in Cr2−δO3 was measured by a tensivolumetric method in the high pO2 range of ≊104 to 104 Pa at 1100°C. The value of δ, or chromium vacancy concentration, was≊9×10−5 mol/mol Cr2O3 in air for Cr2O3 with 99.999% purity. The chemical diffusion coefficient, DT, determined from equilibration data was ≊4.6× cm2·s−1 at 1100°C for pO2= 2.2 ×101 Pa. The self-diffusion coefficient of Cr ions was calculated from and δ and found to be≊1.6×10-17 cm2-s−1, in good agreement with recently measured values.  相似文献   

8.
On heat treatment in air the solubility of MgO or TiO2, in Al23 is too small to detect by lattice parameter shifts. The solubility of MgTiO3 in Al2O3 in air increased to the measured values, expressed as atomic fractions Mg:A1or Ti:A1of0.82 × lo-2, 1.43 × 10-2, and 1.75 × 10-2 at 1250°, 1650°, and 1850°C, respectively. In 1 atm hydrogen the TiO2 solubility expressed as the atomic fraction Ti:A1 is 0.55 × lo-2, 0.75 × 10W2, 1.15 × 10-2, and 1.50 × 10p2 at 1400°, 1500°, 1600°, and 1700°C, respectively. The increased solubility in H2 was attributed to reduction of the titanium ion. The solubility of MgO in A12O3 in vacuum (0.3μ) expressed as the atomic fraction Mg:A1 was measured as 1.10 × loW4, 3.00 × 10"4, 6.80 × 10–4, and 1.40 × 10-3 at 1530°, 1630°, 1730°, and 183O°C, respectively. These contents did not cause an observable change in lattice parameter, but a slight change was observed when MgO was dissolved in A12O3 in a hydrogen atmosphere.  相似文献   

9.
Volatility of 137 Cs and 106Ru from borosilicate glass containing actual high-level waste was measured in an almost closed stainless-steel canister. The temperature dependence of the volatility of 137Cs was close to that obtained in our previous study using 134Cs. The volatility of 106Ru was about one-fifth that of 137Cs at 600° and 800°C. The air contamination by 137Cs and 106Ru in the canister at 400°C was estimated at 1.8 × 102 and 2 × 10 Bq/cm3, respectively, when it was assumed that the glass contained a realistic amount of 137Cs and 106Ru expected in commercial waste glass. These results are useful for predicting safety in a storage facility under operation.  相似文献   

10.
Thermal expansion of the low-temperature form of BaB2O4 (β-BaB2O4) crystal has been measured along the principal crystallographic directions over a temperature range of 9° to 874°C by means of high-temperature X-ray powder diffraction. This crystal belongs to the trigonal system and exhibits strongly anisotropic thermal expansions. The expansion along the c axis is from 12.720 to 13.214 Å (1.2720 to 1.3214 nm), whereas it is from 12.531 to 12.578 Å (1.2531 to 1.2578 nm) along the a axis. The expansions are nonlinear. The coefficients A, B , and C in the expansion formula L t = L 0(1 + At + Bt 2+ Ct 3) are given as follows: a axis, A = 1.535 × 10−7, B = 6.047 × 10−9, C = -1.261 × 10−12; c axis, A = 3.256 × 10−5, B = 1.341 × 10−8, C = -1.954 × 10−12; and cell volume V, A = 3.107 × 10−5, B = 3.406 × 10−8, C = -1.197 × 10−11. Based on α t = (d L t /d t )/ L 0, the thermal expansion coefficients are also given as a function of temperature for the crystallographic axes a , c , and cell volume V.  相似文献   

11.
Third-Harmonic Generation from Some Chalcogenide Glasses   总被引:1,自引:0,他引:1  
Third-order optical nonlinear susceptibilities (χ(3)) of some high-refractive-index chalcogenide glasses were evaluated from third-harmonic generation. Compared with oxide glasses whose χ(3) has been known, χ(3) of the present glasses was higher by an order of magnitude. The addition of selenium drastically increased χ(3). The highest χ(3) was 1.4 × 10–11 esu, being comparable with those of high-χ(3)-organic compounds. Further, χ(3) generally increased with increasing density in the present glasses.  相似文献   

12.
Fluorescence radiation trapping and nonradiative energy losses from the Nd3+4F3/2 state are reported for two widely used commercial phosphate laser glasses (LHG-8 and LG-770). The effects of hydroxyl-group, transition-metal (Cu, Fe, V, Co, Ni, Cr, Mn, and Pt), and rare-earth (Dy, Pr, Sm, and Ce) impurities on the 4F3/2 nonradiative decay rate in these glasses are quantified. Nd concentration quenching effects are reported for doping levels ranging from about 0.5 × 1020 to 8.0 × 1020 ions/cm3. The results are analyzed using the Förster–Dexter theory for dipolar energy transfer. Quenching rates for transition-metal ions correlate with the magnitude of spectral overlap for Nd emission (donor) and the metal ion absorption (acceptor). The nonradiative decay rates due to hydroxyl groups follow Förster–Dexter theory except at low Nd-doping levels (≲2 × 1020 ions/cm3) where the quenching rate becomes independent of the Nd concentration. The data suggest a possible correlation of OH sites with Nd ions in this doping region. The effects of radiation trapping on the fluorescence decay are reported as a function of sample size, shape, and doping level. The results agree well with the theory except for samples with small doping-length products; in these cases, multiple internal reflections from the sample surfaces enhance the trapping effect.  相似文献   

13.
Polycarbosilane-derived SiC fibers (Nicalon) were oxidized at 1773 K under oxygen partial pressures from 102 to 105 Pa. The effect of oxygen partial pressure on the oxidation behavior of the Nicalon fibers was investigated by examining mass change, surface composition, crystal phase, morphology, and tensile strength. The Nicalon fibers were passively oxidized under oxygen partial pressures of >2.5 ×102 Pa and actively oxidized under an oxygen partial pressure of 102 Pa. Under oxygen partial pressures from 2.5 × 102 to 103 Pa, active oxidation occurred at the earliest stage of oxidation, resulting in the formation of both a silica film and a carbon intermediate layer. Although the unoxidized core retained considerable levels of strength under the passive-oxidation condition, fiber strength was lost under the active-oxidation condition.  相似文献   

14.
The thermal conductivities of sintered pellets of ThO2-1.3 wt% U02 were measured at 60°C before and after irradiation. The irradiation temperature was below 156°C, and the exposures varied from 3.1 × 1014 to 4.7 × loL7 fissions/cm3. Each fission fragment damaged a region of 2.2 × 10-16 cm3 with the reduction in conductivity saturating by about 1017 fissions/cm3. Samples having exposures from 1015 to 1016 fissions/cm3 were annealed isothermally at 651 °C or isochronally from 300° to 1200° C to study the annealing of damage. Most of the annealing occurred between 500° and 900°C. The width of this interval plus the slow isothermal annealing suggest that the damage is annealed by a number of single order processes with a spectrum of activation energies from 1.8 to 3.9 eV or, less probably, by a high order process with an activation energy of 3.55 ± 0.4 eV.  相似文献   

15.
The rate of oxygen difusion at 1400°C into pure, deformed, and scandium-doped MgO was measured using a secondary-ion mass spectrometer with 16O- primary ions as sputtering agents to monitor the penetration of 18O into the sample from an 18O-enriched surface layer. In samples doped with scandium, oxygen diffusivity did not differ significantly from the value, 1.3 × 10-15 cm2/s, found for the pure sample. In deformed samples, the diffusion coefficient was 5.8 × 10-15 cm2/s .  相似文献   

16.
New methods of determining the oxygen self-diffusion coefficients (D*o) in oxides have been developed using Raman spectroscopy combined with the 16O–18O exchange technique. From the depth-profiles of the 18O concentration in the 16O–18O exchanged oxides, which was measured by Raman microscope with a spatial resolution of 5 μm, D *o was determined for 2.8 mol% Y2O3-containing tetragonal zirconia polycrystall (the depth-profile method). Thus-obtained results are expressed as D *O,D-P= 1.82(+0.41−0.40) × 10−1·exp{−(139.3 ± 0.2) [kJ/mol]/ RT } [cm2/s] in the temperature range of 700–950°C. We also determined D *o for the same sample from the Raman spectrometric monitoring of the ambient gas during the 16O–18O exchange reaction (the gas-monitoring method). Thus-obtained results are expressed as D *O,G-M= 1.14(+0.05−0.04) × 10−2 exp{−(117.5 ± 0.4) [kJ/mol]/ RT } [cm2/s] in the temperature range of 700–1165°C. The results obtained from the above two different methods virtually agree with each other, indicating that reliable D *o can be obtained by either of these two methods. We demonstrate that Raman spectroscopy is a useful tool for determining D *o in oxides.  相似文献   

17.
Wick debinding was investigated as possible means of binder removal for the Mo-Si-B extrudates. With increased temperature the amount of binder wicked out of the extrudate increased because of a decrease in viscosity, 2513 mPa·s at 75°C down to 7 mPa·s at 250°C. The permeability of the binder in the 1 μm wicking powder was higher than that in the 0.05 μm powder, 3.1 × 10−13 m2 compared to 6.024 × 10−15 m2. The amount of binder removed at a given temperature was considerably lower, which could be attributed to the larger capillary pressure difference between the 0.05 μm wick and the extrudate (1.55 × 105 Pa compared to 1.46 × 104 Pa for the 1 μm powder). Wicking removed approximately 80% of the binder in <10 h at 250°C with no defect formation.  相似文献   

18.
To investigate the effect of reoxidation on the grain-boundary acceptor-state density of reduced barium titanate, n -doped BaTiO3 ceramics are sintered in a reducing atmosphere (2% H2+ 98% N2) and then annealed in oxygen. After annealing at 1150°C for different times, the experimental results show a relationship between temperature-averaged acceptor-state density and annealing time as N s= N so Bt 1/n with n between 2 and 3. An inherent acceptorstate density N so= 4.25 × 1012 cm−2 is obtained with an increase rate B = 4.8 × 1012 cm−2. min−1/3, when n reaches 3. The inherent grain-boundary acceptor states in the reduced n -doped BaTiO3 ceramics are believed not due to adsorbed oxygen ions.  相似文献   

19.
Nucleation and crystal growth rates and properties were studied in a two-stage heat treatment process for Fe2O3-CaO-SiO2 glasses. Glass transition (Tg) and crystallization temperatures (T c ) for the glasses lay between about 612.0° and 710.0°C, and 858.5° and 905.0°C, respectively, and magnetite was the main crystal phase. For a glass of 40Fe2O3. 20CaO·40SiO2 (in wt%) the maximum nucleation rate was (68.6 ± 7) × 106/mm3·s at 700°C, and the maximum crystal growth rate was 9.0 nm/min1/2 at 1000°C. The mean crystal size of the magnetite increased from 30 to 140 nm with variation of nucleation and crystal growth conditions. The glass showed the maxima in saturation magnetization and coercive force, 212.1 × Wb/m2 and 30.8 × 103 A/m, when heat-treated for 4 h at 1000°C and 1050°C, respectively. The variation of the saturation magnetization could be quantitatively interpreted well in terms of the volume fraction of the magnetite, whereas that of the coercive forces could be explained only qualitatively in terms of the particle size of the magnetite. Hysteresis losses showed the maximum value of 1493 W/m3 when heat-treated at 1000°C for 4 h prenucleated at 700°C for 60 min, and increased linearly with increasing heat treatment time under a magnetic field up to 800 × 103 A/m.  相似文献   

20.
The effect of water vapor on the initial sintering kinetics of calcium oxide was studied at 920" to 1123°C. In the range of PHI() from 8 × 10W4 to 487 mm Hg, water vapor increased the rate of sintering. The initial sintering kinetics followed the model for volume diffusion with calculated diffusion coefficients dependent on P H2O, which varied from (P,H2O)1/2 at pressures between 6.1 and 50 mm Hg to (PIH2O)1/2 at pressures between 244 and 487 mm Hg. The activation enthalpy varied from 123 ± 7 kcal/mole when PH2O= 0.04 mm Hg to 103 ± 3.5 kcal/mole at PH2O 243.9 mm Hg.  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