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 31 毫秒
1.
The aim of this study is to explore the effects of Manganese addition and homogenization treatment on the microstructures and mechanical properties of the Al–7Mg–0.15Ti (B535.0) alloy. The optical microscopy, electrical conductivity measurements, transmission electron microscopy, scanning electron microscopy (SEM + EBSD), as well as Rockwell hardness and tensile tests, were exploited for this purpose. The main objectives are to refine the grain size, inhibit grain growth in the annealed state, and enhance the mechanical strength of the alloy. The results show that the addition of manganese to the Al–7Mg–0.15Ti alloys refined the as-cast and recrystallized grains of the alloys. During the homogenization process, Al4Mn high-temperature stable dispersoids were precipitated in the aluminum matrix. After annealing, the Al4Mn particles blocked the movement of grain boundaries during the growth of the recrystallized grains and inhibited grain growth. Consequently, the annealed alloys showed grain refinement and dispersion strengthening. The Al4Mn dispersoids of the alloys with manganese added were smaller and denser after a two-stage homogenization process compared to those that underwent a one-stage homogenization process. By contrast, for the alloys without the addition of manganese, the recrystallized grains showed normal growth after annealing, and different homogenization processes had no significantly different effects.  相似文献   

2.
In order to improve the mechanical properties and to optimize grain refinement of Al-Mg-Si alloy, ECAP processing with an addition of hard particle TiB2 is applied in this work. Mechanical property and microstructural evolution of Al-0.3Mg-7Si+1.5 wt.% TiB2 specimen were investigated by using hardness-testing, optical micrograph observation and electron-backscattering diffraction (EBSD). ECAP processing was done through BA route for 4 passes at room temperature. Hardness test results show that the ECAP process doubled the hardness of the specimen compared to annealed specimen, and from EBSD/OIM analysis, the ECAP processing refined grains from an average grain size of 35 μm to 0.79 μm and led to producing grains having high misorientation angle (≥15).  相似文献   

3.
Quantum chemical DFT calculations at the BP86/TZ2P level have been carried out for the complex [HSi(SiH2NH)3Ti–Co(CO)4], which is a model for the experimentally observed compound [MeSi{SiMe2N(4-MeC6H4)}3Ti–Co(CO)4] and for the series of model systems [(H2N)3M–M′(CO)4] (M = Ti, Zr, Hf; M′ = Co, Rh, Ir). The Ti–Co bond in [HSi(SiH2NH)3Ti–Co(CO)4] has a theoretically predicted BDE of D e = 59.3 kcal/mol. The bonding analysis suggests that the titanium atom carries a large positive charge, while the cobalt atom is nearly neutral. The covalent and electrostatic contributions to the Ti–Co attraction have similar strength. The Ti–Co bond can be classified as a polar single bond, which has only little π contribution. Calculations of the model compound (H2N)3Ti–Co(CO)4 show that the rotation of the amino groups has a very large influence on the length and on the strength of the Ti–Co bond. The M–M′ bond in the series [(H2N)3M–M′(CO)4] becomes clearly stronger with Ti < Zr < Hf, while the differences between the bond strengths due to change of the atoms M′ are much smaller. The strongest M–M′ bond is predicted for [(H2N)3Hf–Ir(CO)4].  相似文献   

4.
Oil palm empty fruit bunch (EFB) is abundantly available in Malaysia and it is a potential source of xylose for the production of high-value added products. This study aimed to optimize the hydrolysis of EFB using dilute sulfuric acid (H2SO4) and phosphoric acid (H3PO4) via response surface methodology for maximum xylose recovery. Hydrolysis was carried out in an autoclave. An optimum xylose yield of 91.2 % was obtained at 116 °C using 2.0 % (v/v) H2SO4, a solid/liquid ratio of 1:5 and a hydrolysis time of 20 min. A lower optimum xylose yield of 24.0 % was observed for dilute H3PO4 hydrolysis at 116 °C using 2.4 % (v/v) H3PO4, a solid/liquid ratio of 1:5 and a hydrolysis time of 20 min. The optimized hydrolysis conditions suggested that EFB hydrolysis by H2SO4 resulted in a higher xylose yield at a lower acid concentration as compared to H3PO4.  相似文献   

5.
Calculations of the vibrational—rotational product state population distributions and differential cross sections for the chemical reaction H + H2(v ? 2, j = 0) → H2(v′ ? 2, j′, mj) + H have been carried out on the Porter—Karplus potential energy surface. The vibrationally-adiabatic-distorted-wave (VADW) method has been used. The relative rotational product distributions, differential cross sections and the helicity mj, dependences of these quantities for the v = 0 reaction agree well with accurate close-coupling results. The absolute integral cross sections are considerably smaller than the accurate quantum values, however. The calculations for the v = 1 reaction agree with the findings of previous sudden quantum, limited close-coupling and quasiclassical theoretical studies and experiments that product H2(v′ = 1) is more likely to be produced than H2(v′ = 0). For the reaction with v = 2, it is found that at high translational energies product H2(v′ = 2) is favoured over H2(v′ = 1) or H2(v′ = 0). The VADW differential cross sections for the v = 1 and v = 2 reactions have a similar shape to those of the v = 0 reaction, with backward peaking when summed over all mj states. The relative rotational distributions for the v = 2, j = 0 → v′ = 2, j and v = 1, j = 0 → v′ = 1, j reactions are also similar to those obtained for the v = 0, j = 0 → v′ = 0, j reaction, with low rotational excitation.  相似文献   

6.
Micron-size poly(methyl methacrylate) (PMMA) particles having a narrow particle size distribution were prepared by seeded dispersion polymerization of methyl methacrylate (MMA) using submicron PMMA particles as seed. The processes of particle aggregation and nucleation were controlled by the initial seed size, initial seed number, and initiator concentration, determining the formation of the mature particles and the number (N (final)) and size of the final particles. It was found that N (final) was equal to the number of particles produced in the absence of seed (N (ab initio)) when the initial number of seed particles (N (initial)) was less than N (ab initio). When N (initial) was greater than N (ab initio), N (final) was equal to k?×?N (initial), where the value of k was a function of seed size and initiator concentration. k increased with seed size and was less than 1 at high initiator concentrations (0.52 and 1.00 %), while at low initiator concentrations (0.23 and 0.30 %), a maximum value of k was found for a 198 nm seed size. k could be greater than unity in some cases.  相似文献   

7.
Computer-aided cooling curve analysis is a reliable method to characterize the solidification behavior of an alloy. In this study, the effect of Al–5Ti–1B grain refiner on the solidification path, microstructure and macrostructure of a new Al–Zn–Mg–Cu super high-strength aluminum alloy containing high amounts of zinc was investigated using thermal analysis technique. The grain size measurement showed that Al–5Ti–1B reduces the grain size from 1402 to 405 μm. Solidification parameters in the liquidus region were in a good accordance with microstructural results. The addition of 1 mass% of Al–5Ti–1B grain refiner decreased ΔT N from 9.1 to 7.7 °C. It also diminished recalescence undercooling from 1.42 to 0.32 °C. The grain refinement also altered dendritic structure of the alloy from a coarse, elongated and non-uniform to a rosette and more uniform shape. Moreover, the grain refiner resulted in a more uniform distribution of eutectic structure between dendrite arms. Furthermore, the grain refinement enhanced fraction of solid at dendrite coherency point from 21 % for unrefined alloy to 25 % for the alloy containing 1 mass% Al–5Ti–1B. In the same trend, the addition of 1 mass% Al–5Ti–1B reduced the amounts of porosity from 2.3 to 1.8 %.  相似文献   

8.
The kinetics of the oxidation of [Ni(II)(H2L1)](ClO4)2, (H2L1 = 3,8-dimethyl-4,7-diaza-3,7-decadiene-2,9-dione dioxime) and [Ni(II)(HL2)]ClO4, (H2L2 = 3,9-dimethyl-4,8-diaza-3,8-undecadiene-2,10-dione dioxime) by peroxodisulfate anion (PDS) in aqueous media at 298.0 K have been studied. The kinetics of oxidation of both Ni(II) complexes was found to be first order in the complex concentration. The dependence of the pseudo-first-order rate constant, k obs, for both complexes showed first-order dependence on PDS concentration. The kinetics of oxidation of [Ni(II)(H2L1)]2+ complex showed a complex dependence on [H+] over the pH range of 4.98–7.50, whereas that of [Ni(II)(HL2)]+ is independent of pH over the pH range of 5.02–7.76. The value of k obs, for both complexes, decreased with increasing ionic strength consistent with the involvement of oppositely charged ions in the rate-determining step. The effect of ionic strength is more pronounced for [Ni(II)(H2L1)]2+–PDS reaction than for [Ni(II)(HL2)]+–PDS reaction, confirming the higher charges of the latter.  相似文献   

9.
Pesticides are widely used in rice cultivation, often resulting in detection of their residues in rice grains. So far, no analytical method has been available for the simultaneous determination of most rice pesticides in rice grains. This paper reports the development and validation of such a method for the determination of eight rice pesticides (penoxsulam tricyclazole, propanil, azoxystrobin, molinate, profoxydim, cyhalofop-butyl, deltamethrin) and 3,4-dichloroaniline, the main metabolite of propanil. Pesticide extraction and clean-up was performed by an optimized matrix solid-phase dispersion (MSPD) protocol on neutral alumina (5 g) using acetonitrile as the elution solvent. Samples were analyzed in a high-performance liquid chromatography–diode array detection (HPLC-DAD) system. Pesticide separation was achieved with a mobile phase of acetonitrile/water in a linear elution gradient from 30:70% (v/v) to 100:0% (v/v) in 14 min at a flow rate of 0.8 mL min?1. Method validation was performed by means of linearity, intra-day accuracy, inter-day precision and sensitivity. Linear regression coefficients (R 2) were always above 0.9948. Limits of detection (LOD) and quantification (LOQ) varied from 0.002 to 0.200 mg kg?1 and 0.006 to 0.600 mg kg?1, respectively. Recoveries were investigated at three fortification levels and were found to be acceptable (74–127%) with relative standard deviations (RSD) below 12%. Application of the method for the analysis of five commercial rice grain samples showed that the pesticide levels were below the LOD. Overall, the method developed is suitable for the determination of residues of most rice pesticides in rice grains at levels below the established MRLs.  相似文献   

10.
Aromatic hydroxylation of mesitylene, phenol and anisole (ArH) with peroxymonosphosphoric acid (H3PO5) in acetonitrile has been studied. H3PO5 is shown to be an effective reagent for aromatic hydroxylation, the reactivity being comparable to that with CF3CO3H. Mesitylene gives mesitol (over 70%). The hydroxylation with H3PO5 is ca 100 fold faster than that with MeCO3H or PhCO3H. The rate equation is: v = k2[ArH][H3PO5] instead of our previous one. The oxidation is catalyzed by H2SO4, giving a linear plot of log k2 vs H0 with a slope of 1.26 for phenol and 1.17 for mesitylene.  相似文献   

11.
The energy spectra of electrons released in thermal energy (≈ 50 meV) ionizing collisions of He*(21 S, 23 S) with H2 have been measured with high resolution and low background. Based on a detailed data analysis, we report accurate H 2 + (v′) vibrational populationsP(v′) for both He*(21 S)+H2(v′=0–10) and He*(23 S)+H2(v′=0–15) and the spectral shapeS(ε) for the individual vibrational peaks. The vibrational populationsP(v′) are quite similar to the Franck-Condon factorsf v ′0 for unperturbed H2(v″=0)→H 2 + (v′) transitions, but, more in detail, the ratiosP(v′)/f v ′0 show a characteristically differentv′-dependence for He*(23 S), He*(21 S), and HeIα(58.4 nm) ionization. The vibrational level separations in the He*(21 S, 23 S)+H2 spectra agree with those in the HeI photoelectron spectrum to within 1–2 meV. The spectral shapesS(ε) are characteristically different for He*(21 S)+H2 and He*(23 S)+H2, reflecting the respective differences in the entrance channel potentials, as determined previously in ab initio calculations and from scattering experiments.  相似文献   

12.
The silsesquioxane [((C6H11)7Si7O9)(OH)3] (LH3) was reacted with [M(C5H5)2Cl2] (M = Ti, Zr, Hf) and with [Ti(C5H5)Cl3]. The reaction with [Ti(C5H5)Cl3] produced [Ti(C5H5)L], whereas the reaction with [Ti(C5H5)2Cl2] produced a mixture of [Ti(C5H5)L]n. (n = 1, 2) as determined by NMR spectroscopy. Only [Ti(C5H5)L] could be isolated from the mixture. The reaction with [M(C5H5)2Cl2] (M = Zr, Hf) produced oligomeric species which contained no cyclopentadienyl ligands and which were formulated as containing trimeric [M3L4Cl] anions on the basis of analytical and spectroscopic data.  相似文献   

13.
The OPAL research reactor in Australia has been used to determine k 0 values for 134mCs, 134Cs, 192Ir and 194Ir. Values for 24Na have also been measured for quality control. The neutron flux at the irradiation positions was very highly thermalised (f > 2,000), resulting in almost negligible activation by epithermal neutrons. As a consequence, the contribution to the total uncertainty of the k 0 values from epithermal-related factors such as Q 0 and $ \bar{E}_{\text{r}} $ was very small. The measured caesium k 0 values have been compared with the library values as well as with recent measurements by St Pierre et al. and Farina Arboccò et al. While there are k 0 values for 194Ir in the library, no 192Ir values have been measured previously. Despite 192Ir having a higher sensitivity than 194Ir, k 0 values were not measured during the establishment of the k 0-method because the nuclear data available at the time indicated that the activation cross-section of 191Ir deviated significantly from 1/v behaviour (g(T n ) ≠ 1), which would result in unacceptable errors if k 0 analysis were to be carried out using the Høgdahl convention. However later nuclear data compilations showed that 191Ir has better 1/v behaviour than previously reported, making it suitable for k 0 analysis using the Høgdahl convention. For completeness, k 0 values have been determined using both the Høgdahl and modified-Westcott conventions and these have been compared with library (194Ir) and calculated values.  相似文献   

14.
To avoid the detection of small fragmentation products of γ-hydroxybutyrate (GHB), a liquid chromatography–tandem mass spectrometry GHB quantification method in human serum supported by adduct formation was developed and validated. The continuous infusion of GHB/GHB-D6 made the identification of two adducts possible and GHB/GHB-D6 sodium acetate adduct fragmentation was used as target mass transition. A Luna 5 μm C18 (2) 100 A, 150 mm?×?2 mm analytical column and elution with a programmed flow of the mobile phase consisting of 10 % A (H2O/methanol = 95/5, v/v) and 90 % B (H2O/methanol = 3/97, v/v), both with 10 mM ammonium acetate and 0.1 % acetic acid (pH?=?3.2), were used. Protein precipitation with 1 mL of the mobile phase B was used as the sample preparation. The calculated limit of detection/quantification was 1 μg/mL. The presented study shows that the fragmentation of GHB sodium acetate adducts is an effective way of quantification of this small molecule and is an interesting alternative to other methods based on the detection of ions smaller than 85 Da. This fact together with the short analysis time of 3 min and the fast sample preparation make this method very attractive for forensic/clinical application.  相似文献   

15.
A kinetic analysis of the oxidation of semicarbazide (SEM) by the single-electron oxidant [IrCl6]2? has been carried out by stopped-flow spectrometric techniques. The reaction proved to be first order each in [IrCl6 2?] and [SEM]tot, leading to overall second-order kinetics. The variation in the observed second-order rate constant k′ with pH was explored over the pH range of 0–7.11. Spectrophotometric titration revealed a stoichiometry of Δ[IrCl6 2?]/Δ[SEM]tot = 4:1 for the redox reaction. On the basis of the rate law, the redox stoichiometry, and the rapid scan spectra, a reaction mechanism is proposed which involves parallel attacks of [IrCl6]2? on both H2NCONHNH3 + and H2NCONHNH2 as rate-determining steps, followed by several rapid reactions. The rate expression, derived from the reaction mechanism, was utilized to simulate the k′–pH profile yielding a virtually perfect fit and indicating that the reaction path involving H2NCONHNH3 + does not make a significant contribution to the overall rate. The reaction between [IrCl6]2? and H2NCONHNH2 was further studied as a function of both temperature and ionic strength. From the temperature dependence, activation parameters were obtained as: ?H 2 ?  = 34.9 ± 1.5 kJ mol?1 and ?S 2 ?  = ?78 ± 5 J K?1 mol?1. The observed ionic strength dependence suggests that the rate-determining step is between [IrCl6]2? and a neutral species of SEM. Hence, both the temperature and ionic strength dependency studies are in good agreement with the proposed reaction mechanism, in which the rate-determining step involves an outer sphere electron transfer.  相似文献   

16.
A new stability-indicating high-performance liquid chromatographic method has been developed for simultaneous analysis of metformin hydrochloride (MET) and sitagliptin phosphate (SIT) in pharmaceutical dosage forms. Chromatographic separation was achieved on a C8 column. The mobile phase was methanol–water 45:55 % (v/v) containing 0.2 % (w/v) n-heptanesulfonic acid and 0.2 % (v/v) triethylamine; the pH was adjusted to 3.0 with orthophosphoric acid. The flow rate was 1 mL min?1 and the photodiode-array detection wavelength was 267 nm. The linear regression coefficients for metformin and sitagliptin were 0.9998 and 0.9996 in the concentration ranges 50–450, and 10–150 μg mL?1, respectively. The relative standard deviations for intra and inter-day precision were below 1.5 %. The drugs were subjected to a variety of stress conditions—acidic and basic hydrolysis, and oxidative, photolytic, neutral, and thermal degradation. The products obtained from photolytic degradation were similar to those from neutral hydrolytic degradation and different from produced by acidic and basic hydrolysis. The method resulted in detection of 15 degradation products (D1–D15); among these, the structures of D1, D3, D9, and D13 were identified. The respective mass balance for MET and SIT was found to be close to 97.60 and 99.12 %. The specificity of the method is suitable for a stability-indicating assay.  相似文献   

17.
The reductions of several substituted acetophenones using supercritical 2-propanol were carried out to estimate the Hammett's reaction constant (ρ=0.33). Also, the reduction of acetophenone using supercritical deuteriated 2-propanol was carried out to determine the rate-determining step. The kinetic isotope effects were observed in the reduction using 2-deuterio-2-propanol (kH/kD=1.6) and O-deuterio-2-propanol (kH/kD=2.0). These findings suggest that the reaction proceeds via a cyclic transition state between acetophenone and 2-propanol similar to that of the Meerwein-Ponndorf-Verley reduction.  相似文献   

18.
19.
The stabilization of olive recombinant hydroperoxide lyases (rHPLs) was investigated using selected chemical additives. Two rHPLs were studied: HPL full-length and HPL with its chloroplast transit peptide deleted (matured HPL). Both olive rHPLs are relatively stable at 4 °C, and enzyme activity can be preserved (about 100% of the rHPL activities are maintained) during 5 weeks of storage at ?20 or at ?80 °C in the presence of glycerol (10%, v/v). Among the additives used in this study, glycine (2.5% w/v), NaCl (0.5 M), and Na2SO4 (0.25 M) provided the highest activation of HPL full-length activity, while the best matured HPL activity was obtained with Na2SO4 (0.25 M) and NaCl (1 M). Although the inactivation rate constants (k) showed that these additives inactivate both rHPLs, their use is still relevant as they strongly increase HPL activity. Results of C6-aldehyde production assays also showed that glycine, NaCl, and Na2SO4 are appropriate additives and that NaCl appears to be the best additive, at least for hexanal production.  相似文献   

20.
Unusually stable [(tC4H9O)3Ti-Co(CO)4] has been prepared by treating the appropriate carbonylmetallate anion with chloro titanium t-butoxide as well as by protolysis of CH3Ti(OtC4H9)3 with HCo(CO)4. Spectroscopic data indicate that the alkoxide and carbonyl ligands are nonbridging, establishing C3v-symmetry at the cobalt atom.  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