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.
A platinum-lined flowing autocláve facility was used to investigate the solubility behavior of magnetite (Fe3O4) in alkaline sodium phosphate and ammonium hydroxide solutions between 21 and 288°C. Measured iron solubilities were interpreted via a Fe(II)/Fe(III) ion hydroxo-, phosphato-, and ammino-complexing model and thermodynamic functions for these equilibria were obtained from a least-squares analysis of the data. A total of 14 iron ion species were fitted. Complexing equilibria are reported for 8 new species: Fe(OH)(HPO4), Fe(OH)2(HPO4)2–, Fe(OH)3(HPO4)2–, Fe(OH)(NH3)+, Fe(OH)2(PO4)3–, Fe(OH)4(HPO4)3–, Fe(OH)2(H2PO4), and Fe(OH)3(H2PO4)3–. At elevated temperatures, hydrolysis and phosphato complexing tended to stabilize Fe(III) relative to Fe(II), as evidenced by free energy changes fitted to the oxidation reactions.
  相似文献   

2.
To determine the solubility product of PuPO4(cr, hyd.) and the complexation constants of Pu(III) with phosphate and EDTA, the solubility of PuPO4(cr, hyd.) was investigated as a function of: (1) time and pH (varied from 1.0 to 12.0), and at a fixed 0.00032 mol⋅L−1 phosphate concentration; (2) NaH2PO4 concentrations varying from 0.0001 mol⋅L−1 to 1.0 mol⋅L−1 and at a fixed pH of 2.5; (3) time and pH (varied from 1.3 to 13.0) at fixed concentrations of 0.00032 mol⋅L−1 phosphate and 0.0004 mol⋅L−1 or 0.002 mol⋅L−1 Na2H2EDTA; and (4) Na2H2EDTA concentrations varying from 0.00005 mol⋅L−1 to 0.0256 mol⋅L−1 at a fixed 0.00032 mol⋅L−1 phosphate concentration and at pH values of approximately 3.5, 10.6, and 12.6. A combination of solvent extraction and spectrophotometric techniques confirmed that the use of hydroquinone and Na2S2O4 helped maintain the Pu as Pu(III). The solubility data were interpreted using the Pitzer and SIT models, and both provided similar values for the solubility product of PuPO4(cr, hyd.) and for the formation constant of PuEDTA. The log 10 of the solubility product of PuPO4(cr, hyd.) [PuPO4(cr, hyd.) \rightleftarrows\rightleftarrows Pu3++PO43-\mathrm{Pu}^{3+}+\mathrm{PO}_{4}^{3-}] was determined to be −(24.42±0.38). Pitzer modeling showed that phosphate interactions with Pu3+ were extremely weak and did not require any phosphate complexes [e.g., PuPO4(aq), PuH2PO42+\mathrm{PuH}_{2}\mathrm{PO}_{4}^{2+}, Pu(H2PO4)2+\mathrm{Pu(H}_{2}\mathrm{PO}_{4})_{2}^{+}, Pu(H2PO4)3(aq), and Pu(H2PO4)4-\mathrm{Pu(H}_{2}\mathrm{PO}_{4})_{4}^{-}] as proposed in existing literature, to explain the experimental solubility data. SIT modeling, however, required the inclusion of PuH2PO42+\mathrm{PuH}_{2}\mathrm{PO}_{4}^{2+} to explain the data in high NaH2PO4 concentrations; this illustrates the differences one can expect when using these two different chemical models to interpret the data. Of the Pu(III)-EDTA species, only PuEDTA was needed to interpret the experimental data over a large range of pH values (1.3–12.9) and EDTA concentrations (0.00005–0.256 mol⋅L−1). Calculations based on density functional theory support the existence of PuEDTA (with prospective stoichiometry as Pu(OH2)3EDTA) as the chemically and structurally stable species. The log 10 value of the complexation constant for the formation of PuEDTA [ Pu3++EDTA4-\rightleftarrows PuEDTA-\mathrm{Pu}^{3+}+\mathrm{EDTA}^{4-}\rightleftarrows \mathrm{PuEDTA}^{-}] determined in this study is −20.15±0.59. The data also showed that PuHEDTA(aq), Pu(EDTA)45-\mathrm{Pu(EDTA)}_{4}^{5-}, Pu(EDTA)(HEDTA)4−, Pu(EDTA)(H2EDTA)3−, and Pu(EDTA)(H3EDTA)2−, although reported in the literature, have no region of dominance in the experimental range of variables investigated in this study.  相似文献   

3.
The initial steps of nucleation of calciumphosphates in aqueous solution are elucidated by means of quantum / classical molecular mechanics simulations. A special focus is dedicated to the role of the protonation state of the phosphate ion. The crystallization of calciumphosphates including entirely deprotonated phosphate ions is found at much lower pH values than required for finding the (PO4)3? species in water. In such cases the depronation of the hydrogenphosphate ion has to occur during crystal growth. According to our findings, the [Ca2+··(PO4)3?··Ca2+] ion triple is the smallest stable aggregate, which may be expected to contain an entirely deprotonated phosphate ion.  相似文献   

4.
Herein, we describe the growth and morphology of well-defined dyed crystals of KH2PO4 (potassium dihydrogen orthophosphate; KDP) containing organic azo (sunset yellow; SSY) dye in the {1 0 1} & {0 0 1} pyramidal growth sectors. An understanding on selective dye inclusion in various growth sector of host crystal is proposed, which will help in designing novel tailor-made dyed photonic crystals. The structural analysis and the identification of various functional groups present in as grown KDP crystals were carried out using powder XRD, FTIR and Raman studies. Solid state transmittance spectra for dyed KDP crystals displayed three absorption peaks at 230 nm, 311 nm and 477 nm, which were blue shifted for SSY dye in KDP crystal relative to neutral aqueous solution of SSY dye. These blue shifts in the absorption maxima confirm the successful incorporation of sunset yellow dye into the pyramidal growth sectors of dyed KDP crystals. The band around 409 nm in the photoluminescence emission spectrum indicates a violet emission. SSY dye doped KDP crystals showed enhanced dielectric properties and thermal stability as compared to pure KDP crystal. The mechanical strength of the KDP crystals estimated using Vickers microhardness test was found to decrease with the increase in SSY dye doping.  相似文献   

5.
In the work presented here, well‐dispersed ferric giniite microcrystals with controlled sizes and shapes are solvothermally synthesized from ionic‐liquid precursors by using 1‐n‐butyl‐3‐methylimidazolium dihydrogenphosphate ([Bmim][H2PO4]) as phosphate source. The success of this synthesis relies on the concentration and composition of the ionic‐liquid precursors. By adjusting the molar ratios of Fe(NO3)3 ? 9H2O to [Bmim][H2PO4] as well as the composition of ionic‐liquid precursors, we obtained uniform microstructures such as bipyramids exposing {111} facets, plates exposing {001} facets, hollow spheres, tetragonal hexadecahedron exposing {441} and {111} facets, and truncated bipyamids with carved {001} facets. The crystalline structure of the ferric giniite microcrystals is disclosed by various characterization techniques. It was revealed that [Bmim][H2PO4] played an important role in stabilizing the {111} facets of ferric giniite crystals, leading to the different morphologies in the presence of ionic‐liquid precursors with different compositions. Furthermore, since these ferric giniite crystals were characterized by different facets, they could serve as model Fenton‐like catalysts to uncover the correlation between the surface and the catalytic performance for the photodegradation of organic dyes under visible‐light irradiation. Our measurements indicate that the photocatalytic activity of as‐prepared Fenton‐like catalysts is highly dependent on the exposed facets, and the surface area has essentially no obvious effect on the photocatalytic degradation of organic dyes in the present study. It is highly expected that these findings are useful in understanding the photocatalytic activity of Fenton‐like catalysts with different morphologies, and suggest a promising new strategy for crystal‐facet engineering of photocatalysts for wastewater treatment based on heterogeneous Fenton‐like process.  相似文献   

6.
Effects of crystal structure on the electrochemistry of boron-doped high-temperature-high-pressure diamond single crystals grown from an Ni–Fe–C–B melt are studied. On the {111}, {100}, and {311} faces, the linear and nonlinear electrochemical impedance spectra and the electrochemical kinetics in the Fe(CN)6 3_/4_ redox system are measured. The acceptor concentration in the diamond interior adjacent to these faces was determined from the Mott–Schottky plots and the amplitude-demodulation measurements. It varies in the 1018 to 1021 cm–3 range. The difference in the electrochemical behavior of individual crystal faces is primarily attributed to different boron acceptor concentrations in the growth sectors associated with the faces.  相似文献   

7.
A platinum-lined, flowing autoclave facility is used to investigate the solubility/phase behavior of zinc(II) oxide in aqueous sodium phosphate solutions at temperatures between 17 and 287°C. ZnO solubilities are observed to increase continuously with temperature and phosphate concentration. At higher phosphate concentrations, a solid phase transformation to NaZnPO4 is observed. NaZnPO4 solubilities are retrograde with temperature. The measured solubility behavior is examined via a Zn(II) ion hydrolysis/complexing model and thermodynamic functions for the hydrolysis/complexing reaction equilibria are obtained from a least-squares analysis of the data. The existence of two new zinc(II) ion complexes, Zn(OH)2(HPO4)2– and Zn(OH)3(H2PO4)2–, is reported for the first time. A summary of thermochemical properties for species in the systems ZnO–H2O and ZnO–Na2O–P2O5–H2O is also provided.  相似文献   

8.
The solubilities of lanthanum carbonate La2(CO3)3·8H2O in solutionsS 0([H+]=H mol kg–1, [Na+]=(IH) mol kg–1, [ClO 4 ]=I mol kg–1) at various fixed partial pressures of CO2 have been investigated at 25.0 °C. The hydrogen ion molality and the total molality of La(III) ion in equilibrium with the solid phase were determined by e.m.f. and analytical methods, respectively. The stoichiometric solubility constants
  相似文献   

9.
Chemical preparation and crystal structure of Pb2Cs3(P4O12)(PO3)3 are described. The triclinic unit cell has the following dimensions: space group is P1 with Z = 2. The crystal structure has been solved by using 3350 unique reflexions with a final R value 0.048, The main outstanding feature of this compound rests on the coexistence in its atomic arrangement of two kinds of phosphate anions with different degrees of condensation: a tetrameric cyclic one (P4O12)4? and a linear infinite chain (PO3). The unit cell is crossed by two (PO3) chains, related by centrosymmetry, running in planes perpendicular to the \documentclass{article}\pagestyle{empty}\begin{document}$ \overrightarrow {\rm c} $\end{document} axis at z ? 0.25 and 0.75. The two crystallographic independent P4O124? ring anions are centrosymmetrical located around (0, 1/2, 0) and (0, 1/2, 1/2) inversion centers, half way between the (PO3) chains. Sr2Cs3(P4O12)(PO3)3 is isotypic with the title compound.  相似文献   

10.
A series of new [NiX(S2P{O-c-Hex}2)(PPh3)](X = Cl, Br, I and NCS)(1)–(4) and [Ni(NCS)(S2P{OR}2)(PPh3)][R =n-Pr (5), i-Pr (6)] complexes has been synthesized and characterized by elemental analyses, f.i.r., i.r., u.v.–vis., 1H-, 13C{1H}- and 31P{1H}-n.m.r. spectra, magnetochemical and conductivity measurements. A single crystal X-ray analysis of [Ni(NCS)(S2P{O-n-Pr}2)(PPh3)](5) reveals the molecular structure of the complex and confirms a square-planar geometry around the central atom of nickel with the NCS anion coordinated via the nitrogen atom.  相似文献   

11.
4‐Amino‐trans‐azobenzene {or 4‐[(E)‐phenyl­diazen­yl]aniline} can form isomeric salts depending on the site of protonation. Both orange bis{4‐[(E)‐phenyl­diazen­yl]anilinium} hydrogen phos­phate, 2C12H12N3+·HPO42−, and purple 4‐[(E)‐phenyl­diazen­yl]­anilinium dihydrogen phosphate phosphoric acid solvate, C12H12N3+·H2PO4·H3PO4, (II), have layered structures formed through O—H⋯O and N—H⋯O hydrogen bonds. Additionally, azobenzene fragments in (I) are assembled through C—H⋯π inter­actions and in (II) through π–π inter­actions. Arguments for the colour difference are tentatively proposed.  相似文献   

12.
A novel phosphate, sodium zinc aluminium bis(phosphate), NaZnAl(PO4)2, was obtained under mild‐temperature hydrothermal conditions at 553 K. The crystal structure has been studied using single‐crystal X‐ray experimental data. The pseudo‐hexagonal phase NaZnAl(PO4)2 crystallizes in the monoclinic space group P21/c. Its unique crystal structure is based on a three‐dimensional (3D) framework built by Zn‐, Al‐ and P‐centred tetrahedra sharing vertices. Channels parallel to the [101] and [01] directions are limited by six‐ and eight‐membered windows, and incorporate Na atoms. The new compound is discussed as a member of the morphotropic series AMM′PO4, where A = Na, K, Rb or NH4, M = Cu, Ni, Co, Fe, Zn or Mg and M′ = Fe, Al or Ga. The title compound is the first Na representative within the series and is characterized by a 3D architecture of tetrahedra populated in an ordered manner by Zn2+, Al3+ and P5+ ions.  相似文献   

13.
The potentiometric method is used to measure the equilibrium potential in the Ti(IV)/Ti(III) system and determine that monophosphate Ti(IV) complexes and Ti3+hydrated complexes dominate in phosphate–perchlorate acid solutions, 4M(H, Na)ClO4, at of 5 × 10–2to 4 × 10–1M. Equations that describe the total electrode reaction are proposed. Decreasing the concentration of free hydrogen ions from 3 to 0.12 M results in the deprotonation of TiO(H2PO4)+complexes and the formation of TiO(HPO4) complexes. Equilibrium constants for reactions of the formation of Ti(IV) monophosphate complexes and the protonation of TiO(HPO4) complex are calculated.  相似文献   

14.
A platinum-lined, flowing autoclave facility is used to investigate the solubility behavior of Cr2O3 and FeCr2O4 in alkaline sodium phosphate, sodium hydroxide, and ammonium hydroxide solutions between 21 and 288°C. Baseline Cr(III) ion solubilities were found to be on the order of 0.1 nmolal, which were enhanced by the formation of anionic hydroxo and phosphato complexes. At temperatures below 51°C, the activity of Cr(III) ions in aqueous solution is controlled by a Cr(OH)3·3H2O solid phase rather than Cr2O3; above 51°C the saturating solid phase is -CrOOH. Measured chromium solubilities were interpreted via a Cr(III) ion hydrolysis/complexing model and thermodynamic functions for the hydrolysis/complexing reaction equilibria were obtained from least-squares analyses of the data. The existence of four new Cr(III) ion complexes is reported: Cr(OH)3(H2PO4), Cr(OH)3(HPO4)2–, Cr(OH)3(PO4)3–, and Cr(OH)4(HPO4)-(H2PO4)4–. The last species is the dominant Cr(III) ion complex in concentrated, alkaline phosphate solutions at elevated temperatures.  相似文献   

15.
The crystal structure of the title bimetallic cyanide‐bridged complex, {K[HoRu(CN)6(H2O)2]·2H2O}n, was determined by means of single‐crystal X‐ray diffraction techniques. The coordination about the central holmium(III) ion is eightfold in a square‐antiprismatic arrangement, while the ruthenium(II) ion is octahedrally coordinated. Channels permeating the crystal lattice contain the potassium cations and two zeolitic water mol­ecules. The HoIII and K atoms lie at sites with mm symmetry and the Ru atom is at a site with 2/m symmetry.  相似文献   

16.
Conductimetric and diaphragm cell techniques have been used to measure diffusion of aqueous potassium phosphate solutions at 25°C from 0.01 to 0.10 mol-dm–3 (M). A significant portion of the aqueous K3PO4 component diffuses as equimolar amounts of potassium hydrogen phosphate and potassium hydroxide produced by hydrolysis: K3PO4+H2O=K2HPO4+KOH. Because OH diffuses more rapidly than HPO 4 2– , the total flow of KOH exceeds the flow of K2HPO4. The extra flow of KOH constitutes coupled transport of a second solute component. Ternary diffusion coefficients that describe interacting flows of K3PO4 and KOH components are reported. At low concentrations where phosphate is strongly hydrolyzed, the molar flux of the KOH component produced by diffusion of K3PO4 is six times larger than the flux of the K3PO4 component. Binary diffusion coefficients for aqueous K2HPO4 solutions are also reported. It is shown that ternary transport coefficients for K3PO4 solutions can be estimated from the properties of binary solutions of K2HPO4 and KOH.  相似文献   

17.
Potentiometric properties of manganese oxides doped with alkali metal ions (Na+, K+, Rb+ and Cs+), which were prepared by heating mixed solutions (starting solution) of each alkali metal and Mn2+ ions, were examined. Electrodes based on mixed phases of Nao44MnO2/Mn2O3 and hollandite KMn8O16/M2O3 found by X-ray powder diffraction (XRD) exhibited Na+- and K+-selective responses with a near-Nernstian slope, respectively, when the molar ratio of alkali metal ion to Mn2+ ion in the starting solution was 0.1. When no alkali metal ions were added in the manganese oxide films, no significant potentiometric response was observed to any alkali metal ions. The selectivity coefficients of these electrodes were = 6.7 × 10–2, = 7.1 × 10–3, < 9 × 10–4 and < 9x 10–4 for the Na0.44MnO2/Mn2O3, and <4 × 10–4 <4x 10–4, =60 × 10–2 ×10–4, < 4 × 10–4, for the KMn8O16/Mn2O3, respectively. Electrodes based on manganese oxides made from mixed solutions of Rb+/Mn2+ and Cs+/Mn2+ also responded to the respective primary ions, that is, Rb+ and Cs+ ions, although XRD patterns for the manganese oxides thus made did not show any peaks except for Mn2O3 (bixbyite); it was concluded in these cases that some amorphous type manganese oxides were formed in the Rb+/Mn2+ and Cs+/Mn2+ systems and they responded to the respective ions. Conditioning of these electrodes in an aerated indifferent electrolyte solution, 0.1M tetramethylammonium nitrate (TMA-NO3), for relatively long time, typically more than 2 hours, was found to be a prerequisite for near-Nernstian response to the respective alkali metal ions. During this electrode conditioning, vacant sites (template) suitable in size for selective uptake of primary ions seemed to be formed by releasing the doped alkali metal ions from the solid phase into the adjacent electrolyte solution accompanying oxidation of the manganese oxide film.  相似文献   

18.
The possibility of formation of various ion clusters for lithium salts LiXF6 (X = As, P) is studied. The dynamic matrix of the clusters in a gas phase is calculated by numerical and analytical differentiation of the full energy of clusters in the MO LCAO approximation by the Hartree-Fock-Roothaan (HFR) method with the aid of program package PC GAMESS. Stable ionic clusters are ion pairs Li+[XF6] with bi- and tridentate cation coordination relative to the octahedral anion, ion triplets [XF6]Li+[XF6] and Li+[XF6]Li+ with bi- and tridentate coordination, and ion dimers { Li+[XF6]}2 with bidentate coordination. Trimers {Li+[XF6]}3 and tetramers {Li+[XF6]}4 in the form of symmetrical ring structures with monodentate coordination are stable only for [AsF6]. For stable ion species, densities of vibrational states and IR spectra are calculated.__________Translated from Elektrokhimiya, Vol. 41, No. 5, 2005, pp. 546–555.Original Russian Text Copyright © 2005 by Popov, Nikiforov, Bushkova, Zhukovskii.  相似文献   

19.
Heavy metals like the actinides possess a high risk potential to the environment not only because of their radiotoxicity but also due to their chemical toxicology. Uranium as one of the major actinide elements has to be considered in particular. Under reducing conditions, tetravalent uranium occurs primarily in the environment. To date, a lack of appropriate analytical techniques that featured sufficient sensitivity made it difficult to study the aqueous phosphate chemistry of uranium(IV) as such complexes show only low solubility. A novel time-resolved laser fluorescence spectroscopy system was set up recently and optimized to do research on uranium(IV). By application of this laser system we could successfully study uranium(IV) phosphate in concentration ranges where no precipitation or formation of colloids occurred. At pH = 1.0, U4+ and one uranium(IV) phosphate complex existed in parallel in aqueous solution. The complex could be identified as [U(H2PO4)]3+. Determination of its corresponding complex formation constant via two different evaluation methods resulted in the finding of (1) logb121 ° = 2 6. 3 7 ±0. 7 6 \log \beta_{121}^{ \circ } = 2 6. 3 7 \pm 0. 7 6 and (2) logb121 ° = 2 6. 4 3 ±0. 2 3 \log \beta_{121}^{ \circ } = 2 6. 4 3 \pm 0. 2 3 . Both values prove that [U(H2PO4)]3+ is a very stable complex in solution under experimental conditions. As they are in very good agreement with each other, the total complex formation constant was determined by means of the weighted average out of (1) and (2). It was calculated to be logb121 ° = 2 6. 4 2 ±0. 2 2 \log \beta_{121}^{ \circ } = 2 6. 4 2 \pm 0. 2 2 .  相似文献   

20.
The solubility of Cd(OH)2(c) was studied in 0.01M NaClO4 solutions, from both the over- and the undersaturation directions, with OH ion concentration ranging from 10–6 to 1.0 mol-L–1, and the equilibration period ranging from 2 to 28 days. Equilibrium Cd concentrations were reached in less than 2 days. The Cd(OH)2(c) solubility showed an amphoteric behavior. In the entire range of OH/H+ investigated, the only dominant aqueous Cd(II) species required to explain the solubility of Cd(OH)2(c) are Cd2+, Cd(OH) 2 0 , and Cd(OH) 4 2– . The logarithms of the thermodynamic equilibrium constants of the Cd(OH)2(c) solubility reactions involving these species, that is, the reactions
  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