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.
Liquid-liquid extraction of cadmium ions from 55-65 w% “black” phosphoric acid from a Nissan H process with diphenyldithio-phosphinic acid (DPP) or dicyclohexyldithiophosphinic acid (DCP) dissolved in an alkane (C12 or C16) at 90°C was investigated. The cadmium concentration could be reduced from 25 ppm to below 2 ppm at a ligand/(Cd+Cu) ratio of 6 (mol/mol). Cd and Cu were removed with a high selectivity. At lower ligand/(Cd+Cu) ratios Cu is removed more selectively than Cd. DPP forms precipitates with Cd, Ni, Zn and Cu, whereas the Cd and Zn complexes of DCP are soluble in the organic phase.

For DCP the equilibrium constant Ke =[CdL2] / ( [Cd] [L¯]2) in 55 w% c.p. H3Po4 (90°C) has a value of 5?105l2/mol2. Ke decreases slightly with decreasing % H3PO4 A fast decomposition of DPP was observed in “black” phosphoric acid at 90°C.  相似文献   

2.
The kinetics of charge-transfer (CT) polymerization of methyl methacrylate (MMA) in the presence of imidazole (Imy) and CCl4 was studied in dimethyl sulfoxide (DMSO) at 60°C. The rate of polymerization (Rp) is sensitive to the [CCl4] up to a concentration of 0.60 mol L−1, but at a higher concentration, it is practically independent of the [CCl4]. When [CCl4] > [Imy], Rp is proportional, to [MMA]1.45±0.15 and [Imy]0.53±0.04 and the average rate constant for the polymerization of MMA is 3.25±0.41 × 10−6 L mol−1 s−1. This article also reports the polymerization of MMA initiated by Imy and CCl4 and accelerated by hexakis (dimethyl sulfoxide) iron (III) perchlorate, [Fe(DMSO)6] (CIO4)3 (A), at 60°C. The presence of Fe(Imy)3+3 in the polymerization system produced well-defined induction periods. The rate constant at 60°C for the interaction of the poly(MMA) radical toward Fe(Imy)3+3 is 7.19 × 104 L mol−1 s−1. A probable reaction mechanism for the polymerization system has been postulated to explain the observed results. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
Whilst the anodic oxidation of (Bu4N) [S4N5] in 0,15 M (Bu4N) [BF4]/CH2Cl2 yielded no detectable radical species, the cathodic reduction at ?40°C and at room temperature leads to epr spectra, which can be assigned to [S4N5]2? and [SN2]? plus a species containing one N atom—possibly NS.2.  相似文献   

4.
This study introduces carbon nanotube buckypaper (CNTBP) into the easily fractured sites of [0°]16 and [0°/90°]4S composite laminates, and comparatively explores how the CNTBP affects the flexural properties of the laminates at 25, ?15, and ?55 °C. Compared to the base [0°]16 and [0°/90°]4S laminates at the same temperature, improvements of the flexural strengths in the order of 4.0–15.3% and 6.5–31.0% are respectively obtained from the corresponding CNTBP‐reinforced [0°]16 and [0°/90°]4S laminates. Importantly, the lower the temperature is, the higher the strength improves. In fact, the CNTBP has little effect on the flexural moduli of the studied laminates, although there is an increasing trend with decreased temperature. Moreover, the introduced CNTBP would significantly change the fracture mechanism of the laminates at low temperature. The present work reveals that the CNTBP exhibits more positive reinforcing capability to the polymer matrix‐based composite laminates at relatively low temperatures.  相似文献   

5.
The polymerization of styrene at 60°C initiated by 2,2′-azobisisobutyronitrile (AIBN) was studied in N,N-dimethylformamide (DMF) in the presence of tris-(bathophenanthroline)iron(III) complex, [Fe(bathophen)3]3+. The complex was prepared in situ by mixing hexakis(N,N-dimethylformamide)iron(III) perchlorate with bathophen-anthroline (systematic IUPAC nomenclature: 4,7-diphenyl-1,10-phenanthroline) in the molar ratio of 1 : 3. The equilibrium constant for was 3.12×103 L3 mol−3. The transfer constant for bathophenanthroline was found to be 0.38 ± 0.01 for the styrene/DMF system at 60°C. Mean velocity constant at 60°C for interaction of polystyryl radical with [Fe(bathophen)3]3+ was 3.73× 104 L mol−1 s−1. © 1996 John Wiley & Sons, Inc.  相似文献   

6.
《Ceramics International》2021,47(19):26869-26876
Converting the γ phase into the α phase completely is necessary in the presintering stage of industrial alumina (Al2O3), which requires high temperature and energy consumption. To reduce the presintering temperature, γ-Al2O3 was activated by oxalic acid. XRD, 27Al-MAS-NMR and TG-DSC were used to characterize the γ - alumina before and after activation, and the phase transformation was studied. The formation temperature of α-Al2O3 decreased to 1029 °C for oxalic acid activated γ-Al2O3, and the α-fraction was 100% for activated γ-Al2O3 at 1300 °C. After oxalic acid activation, the diffraction peak intensity of γ-Al2O3 decreased significantly; the results of 27Al-MAS-NMR suggested that octahedral [AlO6] in γ-Al2O3 was easier than tetrahedral [AlO4] to be attacked by oxalic acid, and the formation of pentavalent [AlO5] with higher reaction activity, which was in favour of the lowering formation temperature of α-Al2O3. The dissolution concentration of Al increased after oxalic acid activation, and the dissolution process was controlled by surface reactions. Oxalic acid mainly attacked the octahedral aluminium in γ-Al2O3 and extracted Al as three complexes of [Al(C2O4)]+, [Al(C2O4)2]- and [Al(C2O4)3]3-. Oxalic acid activated γ - Al2O3 with a lower phase transformation temperature has broad application prospects in the alumina industry.  相似文献   

7.
Two commercial hydroxyapatites were treated in dry ammonia at temperatures between 800 and 1300°C for different times up to 30 h. Nitrogen was incorporated with a maximum content of 3·7 wt% in presence of graphite. X-ray diffraction studies show no important phase transformation by the incorporation of nitrogen up to 1200°C. Treatments at temperatures above 1200°C resulted in the formation of CaO or Ca(OH)2. 31P NMR studies indicate no direct bonding of nitrogen and phosphorous. Infrared spectra show increasing intensities of four new bands at 3250, 2016, 1966 and 700 cm−1 with increasing nitrogen contents, while the OH-bands at 3570 and 630 cm−1 vanish. Taking into account results of the carbon content, XRD, NMR and IR spectra it is suggested that nitrogen enters into the structure as [CN2]2− ion, substituting [OH] groups, and forming cyanamidapatite Ca10(PO4)6(CN2). ©  相似文献   

8.
Diels-Alder Reactions. VII. The Isomerization of 2-Vinylbicyclo[2,2,1]hept-5-ene The isomerization of exo/endo-2-vinylbicyclo[2,2,1]-hept-5-ene in the presence of Na/Al2O3 rapidly gives the thermodynamic equilibrium between exo- and endo-2-vinylbicyclo-[2,2,1]hept-5-ene and cis- and trans-2-ethylidenbicyclo[2,2,1]-hept-5-ene. Far more slowly vinylnortricyclene is formed, which enriches to an equilibrium concentration of about 75% at 60°C. The dependence of equilibrium constants of isomerization on temperature allows the calculation of enthalpies and entropies of isomerization.  相似文献   

9.
H+/Na+ ion-exchange in Ti(HP04) 2.2H20 (γ TiP) was studied. The exchange isotherm, titration and hydrolysis curves at 25°C were obtained. In initially acidic or neutral solutions the 25% of the material exchange capacity was reached. The half-exchanged phase was obtained by additions of stoichiometric amounts of NaOH, with hydrolysis lower than 3% and at acidic equilibrium pH. For additions of NaOH over 4.0 me/g Y-TiP conversions higher than 50% were reached with an increase in the values of the equilibrium pH and the hydrolysis. The evolution of the solid was followed by X-ray diffraction. The substituted phases of samples dried in air at room temperature showed the following hydration degrees: ----H N ----------------------- conversion phase stored over P205 was dehydrated originating H1-5Na0.5 When i t is heated at 80°C-200°C i t transformed i n t o mixtures o f iTFi ( 9. 1 A ) and HNa (10.1 Å ).  相似文献   

10.
Polymerization of methyl methacrylate (MMA) by the charge-transfer complex formed by the interaction of 2,2′-bipyridine (bpy), MMA, and carbon tetrachloride (CCl4) was studied in dimethylsulfoxide (DMSO) at 60°C. The rate of polymerization (Rp) is sensitive to the [CCl4] at low concentration of CCl4, but at a higher concentration it is practically independent of [CCl4]. Rp is proportional to [MMA]1.45±0.04 and [bpy]0.52±0.04 when [CCl4] > [bpy], and the average rate constant, k, at 60°C for the polymerization of MMA was 7.14 ± 0.40 × 10−6 L mol−1s−1. Kinetic studies showed that the polymerization proceeds through free radical intermediates. This article also reports the polymerization of MMA initiated by bpy and CCl4 and accelerated by Lewis acid, hexakis (dimethylsulfoxide)iron(III) perchlorate [Fe(DMSO)6](ClO4)3 at 60°C. The glass transition temperature and molecular weights of the samples were investigated by using differential scanning calorimetry and gel permeation chromatography techniques, respectively. Probable reaction mechanisms are proposed to explain the observed results. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 64: 2097–2103, 1997  相似文献   

11.
Mixed titanates Ba1−xSrxTiO3 were synthesized via calcination of oxalate coprecipitated precursors. On heating there were three thermal event occurred: T<250°C corresponds to the evaporation of trapped water and dehydration, T=250°C–450°C corresponds to the decomposition of oxalate and the formation of an intermediate phase where the composition is close to [Ba1−xSrx]2Ti2O5.CO3 and T=600°C–700°C corresponds to the carbonate decomposition and the formation of Ba1−xSrxTiO3 phase. Powders calcined at T=700°C for 2 h are single phase, have grain size ranges of 0.2–2 μm and elongated morphology. Rietveld refinements of the XRD data showed single phase of perovskite structure in which their tetragonality decreased with increasing concentration of Sr2+ incorporated in Ba2+ site. The transition temperature showed strong correlation with the tetragonality.  相似文献   

12.
Ti3SiC2 phase was synthesized by reactive pyrolysis of three different polycarbosilane, [Si(H)2CH2]m·[(H)Si(Vi)CH2]n·[(H)Si(Me)CH2]p (AHPCS), [Si(CH3)2CH2]x·[Si(H)(CH3)CH2]y (PCS), and [Me(H)SiC≡C]n, filled with metal Ti powder. The pyrolysis was carried out in argon atmosphere between 1200°C and 1400°C. The metal–precursor reactions and phase evolution during the pyrolysis were studied by means of X-ray diffraction and scanning electron microscopy. The results indicated that PCS/Ti system was more beneficial for the synthesis of Ti3SiC2. In addition, the high-purity Ti3SiC2 could be synthesized through the pyrolysis of green compact of the PCS/Ti system with CaF2 at 1400°C.  相似文献   

13.
Visible absorption spectra and the molar conductance curve of NiCl2 dissolved in dimethyl-sulphoxide (DMSO) have been determined at 25°C. The results indicate the formation of the [NiCl(DMSO)5]+ inner-sphere complex followed by the formation of the {NiCl(DMSO)5]+Cl?} outersphere ion-pair as the concentration of NiCl2 increases. The [NiCl2(DMSO)4] inner-sphere complex in pure dimethyl sulphoxide at 25°C is formed to but a small extent, but its relative content increases upon addition of the non-coordinating diluent, chlorobenzene. Dilutions with chlorobenzene also enhance coordination disproportionation producing the [NiCl3 DMSO]? tetrahedral complex. The stability constant of the monochloro complex derived from the conductometric data has the value of 297 (±10)M?2 at 25°C, and the estimated value of the association constant of the [NiCl (DMSO)5]+Cl? complex electrolyte, dominating in moderately concentrated solutions at room temperature, is 7(±3)M?1.  相似文献   

14.
LiNiPO4 (LNP) ceramics were synthesized using a conventional solid-state reaction method and different sintering temperatures. Differential scanning calorimetry (DSC), thermogravimetric (TG) analysis, and X-ray diffraction (XRD) measurements indicated that single- phase olivine (Pnma, No. 62) was formed above 750°C, and dense LNP ceramic with a theoretical density of more than 95% was obtained at 825°C. Raman and far-infrared (IR) vibrational modes were assigned and discussed in detail. The intrinsic dielectric properties of the samples were calculated using the four-parameter semi-quantum (FPSQ) model based on far-IR reflectance spectroscopy and were in good agreement with the measured values. A positive relationship existed between the Raman shift of the υ1 mode (attributed to the symmetric vibration of [PO4]3−) and the corrected permittivity, and the opposite correlation was observed between the quality factor (× f) and the damping of the υs mode as well as the distortion of the [NiO6] octahedra. The optimized microwave dielectric properties of the LNP ceramics sintered at 825°C include an ultralow dielectric constant (5.18) and a good quality factor (24 076 GHz, f = 17.2 GHz).  相似文献   

15.
Abstract

(±)?Syn?dibenzo[a,l]pyrene diol epoxide (DB[a,l]PDE) and (±)?anti?DB[a,l]PDE were reacted with deoxyadenosine (dA) or deoxyguanosine (dG) in dimethylformamide at 100 °C for 30 min. The crude products were purified by reverse phase HPLC under gradient and isocratic conditions. The structure of each adduct was assigned by 1D and 2D NMR spectra and by fast atom bombardment mass spectrometry. Five adducts were isolated from the reaction of (±)?syn?DB[a,l]PDE and dA: syn?DB[a,l]PDE?N6dA?1, syn?DB[a,l]PDE?N6dA?2, syn?DB[a,l]PDE?N6dA?3, syn?DB[a,l]PDE?N6dA?4 and syn?DB[a,l]PDE?N7Ade. Four adducts were isolated from the reaction of (±)?anti?DB[a,l]PDE and dA: anti?DB[a,l]PDE?N6dA?1, anti?DB[a,l]PDE?N6dA?2, anti?DB[a,l]PDE?N6dA?3 and anti?DB[a,l]PDE?N6dA?4. Two adducts were isolated from the reaction of (±)?syn?DB[a,l]PDE and dG: (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N2dG and (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N7Gua. Two adducts were isolated from the reaction of (±)?anti?DB[a,l]PDE and dG: (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N2dG and (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N7Gua.  相似文献   

16.
Polysiloxanes [RSiO1.5]n with R=CH3 (PMS) and C6H5 (PPS), respectively, were transformed to Si–O–C ceramics of variable composition and structure upon pyrolysis in inert atmosphere at 800–1500°C. The electrical conductivities of the Si–O–C ceramics in air were measured at room temperature by using a shielded two point configuration. In situ measurements of the dc-conductivity during the pyrolytic conversion from the polymer to the ceramic phase were carried out up to 1500°C with four point contacted carbon electrodes in inert atmosphere. During polymer-ceramic conversion excess carbon precipitates above 400°C (PPS)–700°C (PMS). At temperatures above 800°C (PPS) and 1400°C (PMS) coagulation and growth of the carbon clusters results in a percolation network formation. While below the percolation threshold electrical conductivity can be described according to Motts mechanism by variable-range-hopping of localized charge carriers, regular electron band conduction due to the instrinsic conductivity of turbostratic carbon (8×10−4 (Ωcm)−1) predominates above. Thus, the in situ measurement of non-linear electrical property changes can be used as a microprobe of high sensivity to detect microstructural transformations during the pyrolysis of preceramic polymers.  相似文献   

17.
The solubility of potassium ferrate (K2FeO4) was measured in aqueous solutions of NaOH and KOH of total concentration 12 M containing various molar ratios of KOH:NaOH in the range 12:0 to 3:9. Several analytical methods were tested for the determination of ferrate concentration. The final method chosen consisted of potentiometric titration of the ferrate sample with an alkaline solution of As2O3. The assumption was made that ferrate dissociates in concentrated KOH solutions predominantly to KFeO4. The solubility constant, S, defined as the product of the molar concentration of the potassium ion, K+, and the ferrate anion, KFeO4, was found to be 0·044 ± 0·006 mol2 dm−6 for 20°C, 0·093 ± 0·004 mol2 dm−6 for 40°C and 0·15 ± 0·09 mol2 dm−6 for 60°C. From these results the heat of dissolution of K2FeO4 was calculated as −14·3 kJ mol−1. At 60°C the enhanced decomposition of the ferrate at the higher temperature led to a greater deviation in solubility values compared with data for either 20°C or 40°C.  相似文献   

18.
The polymerization of caprolactone (ϵ-CL) was initiated by yttrium triisopropoxide {Y(OPir)3}, bimetallic isopropoxide of yttrium and aluminum {Y[Al(OPis)4]3}, yttrium and tin(II) {Y[Sn(OPir)3]3}, and tin(II) and yttrium {Sn[Y(OPir)4]2}, respectively. The polymerization was carried out through coordinative insertion of a monomer into the free metal-oxygen bond. The molecular weight and yield of poly(ϵ-caprolactone) (PCL) were affected drastically by the mol % of the initiator to the monomer (Co/Mo). The results showed that Y(OPir)3 and Sn[Y(OPir)4]2 were more effective than were {Y[Al(OPir)4]3} and {Y[Sn(OPir)3]3} for the polymerization of ϵ-CL. When polymerization was conducted at 5°C using Y(OPir)3 as the initiator or at 10°C using Sn[Y(OPir)4]4 as the initiator, polymers with a molecular weight of 45.9 × 103 or 54.0 × 103 and high yield resulted in 30 or 5 min, respectively. The polymer was characterized by FTIR, H-NMR, and GPC. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 64: 1295–1299, 1997  相似文献   

19.
Red phosphor BaSiF6:Mn4+ has been synthesized by a hydrothermal method at 120°C for 24 h, in which either Si or SiO2 is used as silicon source. HF as weak acid performs a complex agent for the formation of anion groups [SiF6]2? and [MnF6]2?. The luminescence properties of undoped BaSiF6 have been firstly observed. The dependence of luminescence intensities of BaSiF6:Mn4+ on the concentrations of HF and KMnO4 in precursory solution has been investigated. The as‐prepared BaSiF6:Mn4+ exhibits high chemical stability even in deionized water.  相似文献   

20.
The effects of Fe2O3 on phase evolution, density, microstructural development, and mechanical properties of mullite ceramics from kaolin and alumina were systematically studied. X-ray diffraction results suggested that the ceramics consisted of mullite, sillimanite, and corundum, in the sintering range of 1450°C–1580°C. However, as the sintering was raised to 1580°C, mullite is the main phase with a content of 94%, and the corundum phase content is 5.9%. Simultaneously, high-temperature sintering had a positive effect on the densification of the mullite ceramics, where both the bulk density and flexural strength could be optimized by adjusting the content of Fe2O3. It was found that 6 wt% Fe2O3 was optimal for the formation of rod-shaped mullite after sintering at 1550°C for 3 h. The sample's maximum bulk density was 2.84 g/cm3, with a flexural strength of 112 MPa. Meanwhile, rod-shaped mullite grains with an aspect ratio of ~9 were formed. As a result, a dense network structure was developed, thus leading to mullite ceramics with excellent mechanical properties. The effect of Fe2O3 on the properties might be attributed to the fact that Al3+ ions in the [AlO6] octahedron were replaced by Fe3+ ions, resulting in lattice distortion.  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