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 autoclave facility is used to investigate the solubility behavior of copper(II) oxide (CuO) in aqueous sodium phosphate solutions at temperatures between 19 and 262°C. Copper solubilities are observed to increase continuously with temperature and phosphate concentration. The measured solubility behavior is examined via a Cu(II) ion hydrolysis/complexing model and thermodynamic functions for the hydrolysis/complexing reactions are obtained from a leastsquares analysis of the data. Altogether, thermochemical properties are established for five anionic complexes: Cu(OH) 3 , Cu(OH) 4 2– , Cu(OH) 2 (HPO 4 ) 2– , Cu(OH) 3 (H 2 PO 4 ) 2– , and Cu(OH) 2 (PO 4 ) 3– . Precise thermochemical parameters are also derived for the Cu(OH)+ hydroxocomplex based on CuO solubility behavior previously observed (Ref. 3) for pure water at elevated temperatures. The relative ease of Cu(II) ion hydrolysis is such that Cu(OH) 3 species become the preferred hydroxocomplex for pH9.4.Prepared for presentation at the Fourth International Symposium on solubility Phenomena, Rensselaer Polytechnic Institute, August 1990.  相似文献   

2.
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   

3.
A platinum-lined, flowing autoclave facility was used to investigate the solubility behavior of titanium dioxide (TiO2) in aqueous sodium phosphate, sodium hydroxide and ammonium hydroxide solutions between 17 and 288°. Baseline Ti(IV) solubilities were found to be on the order of one nanomolal, which were enhanced by the formation of anionic hydroxo- and phosphato-complexes. The measured solubility behavior was examined via a titanium(IV) ion hydrolysis/complexing reaction equilibria were obtained from a least squares analysis of the data. The existence of three new Ti(IV) ion complexes is reported for the first time: Ti(OH)4(HPO4)2–, Ti(OH)5(H2PO4)2– and Ti(OH)5(HPO4)3–. The triply-charged anionic complex was the dominant Ti(IV) species in concentrated, alkaline phosphate solutions at elevated temperatures. This complex is expected to exhibit C.N.=4 (i.e., Ti(OH)2OPO 4 3– ). A summary of thermochemical properties for species in the systems TiO2-H2O and TiO2-P2O5-H2O is also provided.  相似文献   

4.
A platinum-lined, flowing autoclave facility was used to investigate the solubility/phase behavior of nickel oxide (NiO) in aqueous sodium phosphate solutions between 290 and 560 K. A layer of hydrous nickel oxide was concluded to exist on the nickel oxide surface below 468 K; only at higher temperatures did the anhydrous nickel oxide phase control the nickel ion solubility behavior. The measured solubility behavior was examined via a nickel(II) ion hydrolysis/complexing model and thermodynamic functions for the hydrolysis/complexing reaction equilibria were obtained from a least-squares analysis of the data. The existence of two new nickel ion complexes are reported for the first time: Ni(OH)2(HPO4)= and Ni(OH)3(H2PO4)=. The positive entropy change associated with the formation of Ni(OH)3(H2PO4)= leads to its dominance in alkaline phosphate solutions at elevated temperatures.  相似文献   

5.
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   

6.
Raman spectra have been measured for aqueous ZnSO4 solutions under hydrothermal conditions at steam saturation to 244°C; solubility has been recorded as a function of temperature from 25 to 256°C. The high-temperature Raman spectra contained two polarized bands, which suggest that a second sulfato complex, possibly bidentate, is formed in solution, in addition to the 1:1 zinc(II) sulfato complex, which is the only ion pair identified at lower temperatures. Under hydrothermal conditions, it was possible to observe the hydrolysis of the zinc(II) aquo ion by measuring the relative intensity of bands due to SO 4 2– and HSO 4 according to the equilibrium reaction Zn(OH2)6]2+ + SO 4 2– [Zn(OH2)5OH]+ + HSO 4 The precipitate in equilibrium with the solution at 210°C could be characterized as ZnSO4 · H2O (gunningite) by x-ray diffraction (XRD) and Raman and infrared spectroscopy. At 244°C the equilibrium precipitate could be identified as ZnSO4 (zincosite).  相似文献   

7.
An aqueous thermodynamic model is developed which accurately describes the effects of high base concentration on the complexation of Ni2+ by ethylenedinitrilotetraacetic acid (EDTA). The model is primarily developed from an extensive dataset on the solubility of Ni(OH)2(cr) in the presence of EDTA and in the presence and absence of Ca2 + as the competing metal ion. The solubility data for Ni(OH)2(cr) were obtained in solutions ranging in NaOH concentration from 0.01 to 11.6 mol-kg–1, and in Ca2 + concentrations extending to saturation with respect to portlandite, Ca(OH)2. Owing to the inert nature of the Ni-EDTA complexation reactions, solubility experiments were approached from both the oversaturation and undersaturation direction and over time frames extending to 413 days. The final aqueous thermodynamic model is based upon the equations of Pitzer, accurately predicts the observed solubilities to concentrations as high as 11.6 mol-kg–1 NaOH, and is consistent with UV–Vis spectroscopic studies of the complexes in solution.  相似文献   

8.
Chromium(III)-phosphate reactions are expected to be important in managing high-level radioactive wastes stored in tanks at many DOE sites. Extensive studies on the solubility of amorphous Cr(III) solids in a wide range of pH (2.8–14) and phosphate concentrations (10–4 to 1.0 m) at room temperature (22±2)°C were carried out to obtain reliable thermodynamic data for important Cr(III)-phosphate reactions. A combination of techniques (XRD, XANES, EXAFS, Raman spectroscopy, total chemical composition, and thermodynamic analyses of solubility data) was used to characterize solid and aqueous species. Contrary to the data recently reported in the literature,(1) only a limited number of aqueous species [Cr(OH)3H2PO4, Cr(OH)3(H2PO4)2–2), and Cr(OH)3HPO2–4] with up to about four orders of magnitude lower values for the formation constants of these species are required to explain Cr(III)-phosphate reactions in a wide range of pH and phosphate concentrations. The log Ko values of reactions involving these species [Cr(OH)3(aq)+H2PO4⇌Cr(OH)3H2PO4; Cr(OH)3(aq)+2H2PO4⇌Cr(OH)3(H2PO4)2–2; Cr(OH)3(aq)+HPO2–4⇌Cr(OH)3HPO2–4] were found to be 2.78±0.3, 3.48±0.3, and 1.97±0.3, respectively.  相似文献   

9.
The stability constants of zirconium(IV) hydrolysis species have been measured at 15, 25, and 35 °C [in 1.0 mol-dm–3 (H,Na)ClO4] using both potentiometry and solvent extraction. In addition, the solubility of [Zr(OH)4(am)] has been investigated in a 1 mol-dm–3 (Na,H)(ClO4,OH) medium at 25 °C over a wide range of –log [H+] (0-15). The results indicate the presence of the monomeric species Zr(OH)3+, Zr(OH)2 2+, Zr(OH)3 +, and Zr(OH)4 0(aq) as well as the polymeric species Zr4(OH)8 8+ and Zr2(OH)6 2+. The solvent extraction measurements required the use of acetylacetone. As such, the stability constants of zirconium(IV) with acetylacetone were also measured using solvent extraction. All stability constants were found to be linear functions of the reciprocal of temperature (in kelvin) indicating that H o and S o are both independent of temperature (over the temperature range examined in the study). The results of the solubility experiments have shown four distinctly different solubility regions. In strongly acidic solutions, the solubility is controlled by the formation of polynuclear hydrolysis species in solution whereas in less acidic solution the formation of mononuclear hydrolysis species becomes dominant. The largest portion of the solubility curve is controlled by equilibrium with aqueous Zr(OH)4 0(aq) where the solubility is independent of the proton concentration. In alkaline solutions, the solubility increases due to formation of the zirconate ion. The middle region was used to determine the solubility constant (log *K s10) of Zr(OH)4(s). From the data in the alkaline region, a value of the stability of the zirconate ion has been determined. This is the first time that the possible evidence for the zirconate ion has been identified in aqueous solution that has previously been found only in the solid phase.  相似文献   

10.
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   

11.
Zinc hydroxide chloride particles were synthesized by hydrolysis of ZnCl2 solutions dissolving AlCl3 at different atomic Al/Zn ratios from 0 to 1.0 and characterized by various techniques. Increasing Al/Zn ratio changed the crystal phases of the products as ZnO→ZnO+ZHC (Zn5(OH)8Cl2·H2O)→ZHC→LDH (layered double hydroxides, Zn-Al-Cl) and the particle morphology as agglomerates (ZnO)→fine particles (ZnO)→plates (ZHC)+rods (ZnO)→plates (ZHC)→plates (LDH). The atomic Cl/Zn ratios of LDH particles formed at Al/Zn?0.3 were ca. 0.3 despite the increase of Al/Zn ratio, being due to the intercalation of CO32− into the LDH crystal. The OH content of LDH estimated by TG was reduced by the deprotonation of OH to counteract the excess positive charge produced by replacing Zn(II) with Al(III). ZHC exhibited a high adsorption selectivity of H2O.  相似文献   

12.
The equilibrium potential of saturated zinc amalgam is studied as a function of concentration of free ethylenediamine molecules, [en], in the region [en] 0.001–1 M in solutions of pH 9.5, 10.5, and 11.5. At the concentration of zinc(II) ions 2 × 10–3 M and [en] = 1 M only simple trisethylenediamine complexes of zinc(II) form in all the solutions. At smaller [en] and pH 9.5 and 10.5, complexes Zn(en)2 2+ and Zn(en)2OH+ are also present; these are complemented at pH 11.5 by Zn(en)2(OH)2 at [en] 0.005–0.1 M. Stability constants for these complexes are calculated.  相似文献   

13.
Herein, we studied the experimental and theoretical foundations of the process of zinc(II) and cadmium(II) complexation with 2-hydroxido-nonahydrido-closo-decaborate(2−) anion [2-B10H9(OH)]2− in the presence of azaheterocyclic ligands L (L=2,2′-bipyridyl (bipy), 1,10-phenanthroline (phen), and 2,2′-bipyridylamine (bpa)), which can be used as model system for obtaining complexes with the required composition and structure. The first examples of mixed-ligand Zn(II) and Cd(II) complexes with [2-B10H9(OH)]2− coordinated by the metal atom were isolated selectively. The structures of zinc(II) complexes [Zn(bipy)2(2-B10H9(OH)-κ2H1,O)] ⋅ 2CH3CN ( 1 ⋅ 2CH3CN) and [Zn(phen)2(2-B10H9(OH)-κ2H9,O)] ⋅ 2CH3CN ( 2 ⋅ 2CH3CN), as well as two cadmium(II) bond isomers [Cd(bipy)2(2-B10H9(OH)-κ2H1,O)] ( 4 a ) and [Cd(bipy)2(2-B10H9(OH)-κ2H9,H10)] ( 4 b ) bound into a dimeric pair in the complex [Cd(bipy)2(2-B10H9(OH))] ( 4 ), and cadmium(II) complex [Cd(bpa)2(2-B10H9(OH)-κ2H7,H10)] ( 7 ) were solved by single-crystal X-ray diffraction (XRD). Density functional theory (DFT) calculations show that for cadmium(II) the formation of both multicenter BH−Cd−HB and BO(H)−Cd−HB bonds is equally probable. The affinity of zinc(II) for oxygen leads to preferential formation of complexes via BO(H)−Zn−HB bonds than BH−Zn−HB bonds. The M−B(H) bonding was found to be presumably electrostatic in nature, which could be the reason of topological isomerism of zinc(II) and cadmium(II) decaborates.  相似文献   

14.
The objectives of this study were to address uncertainties in the solubility product of (UO2)3(PO4)2⋅4H2O(c) and in the phosphate complexes of U(VI), and more importantly to develop needed thermodynamic data for the Pu(VI)-phosphate system in order to ascertain the extent to which U(VI) and Pu(VI) behave in an analogous fashion. Thus studies were conducted on (UO2)3(PO4)2⋅4H2O(c) and (PuO2)3(PO4)2⋅4H2O(am) solubilities for long-equilibration periods (up to 870 days) in a wide range of pH values (2.5 to 10.5) at fixed phosphate concentrations of 0.001 and 0.01 M, and in a range of phosphate concentrations (0.0001–1.0 M) at fixed pH values of about 3.5. A combination of techniques (XRD, DTA/TG, XAS, and thermodynamic analyses) was used to characterize the reaction products. The U(VI)-phosphate data for the most part agree closely with thermodynamic data presented in Guillaumont et al.,(1) although we cannot verify the existence of several U(VI) hydrolyses and phosphate species and we find the reported value for formation constant of UO2PO4 is in error by more than two orders of magnitude. A comprehensive thermodynamic model for (PuO2)3(PO4)2⋅4H2O(am) solubility in the H+-Na+-OH-Cl-H2PO4-HPO2−4-PO3−4-H2O system, previously unavailable, is presented and the data shows that the U(VI)-phosphate system is an excellent analog for the Pu(VI)-phosphate system.  相似文献   

15.
An inert, flowing autoclave facility was used to investigate the solubility behavior of α-PbO (litharge, tetragonal) in aqueous solutions of morpholine, ammonia and sodium hydroxide between 38 and 260 C. Lead solubilities increased from about 0.4 mmol-kg−1 at 38 C to about 4.5 mmol-kg−1 at 260 C and were relatively insensitive to the concentration and identity of the reagent used to control the pH. The measured lead solubilities were interpreted using a Pb(II) ion hydroxocomplexing model and thermodynamic functions for these equilibria were obtained from a least-squares analysis of the data. A consistent set of thermodynamic properties for the species Pb(OH)+, Pb(OH)2(aq) and Pb(OH)3 was obtained that will permit accurate lead oxide solubility calculations to be made over broad ranges of temperature and alkalinity.  相似文献   

16.
The solubility product of the solid hydroxides and the first hydrolysis constants of trivalent ions of lanthanum, praseodymium and lutetium, were determined in 2 M NaClO4(aq) and 2 M NaCl(aq) at 303 K, where M denotes the concentration in mol-L−1. Solubility diagrams (pLn(aq)−pCH) were measured by means of a radiochemical method. The pCH borderlines of precipitation and the solubility products were determined from these diagrams. The fitting of the solubility equation with the experimental values from the pLn(aq)−pCH diagrams also allowed the calculation of the first hydrolysis constants and the solubility products. In separate experiments, the stability constants for the first monohydroxide species were determined by means of potentiometric pH titrations, where the data were treated with both the program SUPERQUAD and by fitting of the results to the mean ligand number equation. Values of the log10 < eqid20 > 1,Cl constants for the LnCl2+ species were also calculated at 2 M ionic strength and 303 K, using the hydrolysis constants obtained in both perchlorate and chloride media. The quantitative effects of chloride ions on the hydrolysis reactions and solubilities were determined for these three rare-earths spanning the lanthanide series.  相似文献   

17.
A Raman spectral study of 14 solutions of varying bromide to zinc ratios was conducted up to 300°C and 9 MPa. The tetra-, tri-, di- as well as the mono-bromozinc complexes were identified. The signal from the ZnBr+ complex increased in intensity as temperature increased, for solutions of low bromide- to-zinc ratios. The ZnBr 4 2– species was favored at higher Br/Zn ratios, and higher temperatures favored the formation of the species ZnBr2 and ZnBr+ at the expense of ZnBr 4 2– and ZnBr 3 . Although solvated water is probably present in these zinc-bromo complexes, we found no evidence of O–Zn vibrations other than for Zn(H2O) 6 2+ . However, spectra of successive dilutions of solutions with high bromide to zinc ratios show a relative change in species populations thereby suggesting that water activity plays a decisive role in complex formation. For the first time trifluoromethanesulfonic acid (HTFMS) has been used as an internal standard in Raman spectroscopy. This permitted quantitative measurement of stepwise stability constants.  相似文献   

18.
Isosaccharinate (ISA), a degradation product of cellulose codisposed in low-level nuclear wastes, is expected to be one of the dominant complexing ligands for radionuclides, especially tetravalent actinides. This paper presents a comprehensive thermodynamic model for isosaccharinate reactions with Ca(II) and Np(IV). The model is valid for a wide range of pH values (2–14), ISA concentrations (ranging up to 0.1 m), and ionic strengths (ranging up to 6.54 m), and is based on (1) NMR investigations of HISA(aq) (-D-isosaccharinic acid) and ISL(aq) [dehydration product of HISA(aq)], and the solubility of Ca(ISA)2(c) as a function of pH and concentrations of Ca and ISA; (2) NpO2(am) solubility in a wide range of pH values (2–14) and total ISA concentrations of 0.0016 and 0.008 m and at fixed pH values of approximately 5 and 12 with total ISA concentrations ranging from 0.0001 to 0.1 m; and (3) solvent extraction of Np-ISA solutions, containing fixed NaClO4 concentrations ranging from 0.103 to 6.54 m and at fixed pC H+ values ranging from 1.5 to 1.9, with dibenzoylmethane. Pitzer's ion-interaction approach was used to interpret the data. The different aqueous species required to explain these data included HISA(aq), ISL(aq), Ca(ISA)+, Np(OH)3(ISA)(aq), Np(OH)3(ISA)2 , Np(OH)4(ISA), and Np(OH)4(ISA)2 2–. The values of equilibrium constants for reactions involving these species and determined from these data provided close agreements between the observed and predicted concentrations in all of the systems investigated in this study and those reported previously.  相似文献   

19.
The main objective of this study was to develop a thermodynamic model for predicting Cr(III) behavior in concentrated NaOH and in mixed NaOH–NaNO3 solutions for application to developing effective caustic leaching strategies for high-level nuclear waste sludges. To meet this objective, the solubility of Cr(OH)3(am) was measured in 0.003 to 10.5 m NaOH, 3.0 m NaOH with NaNO3 varying from 0.1 to 7.5 m, and 4.6 m NaNO3 with NaOH varying from 0.1 to 3.5 m at room temperature (22 ± 2°C). A combination of techniques, X-ray absorption spectroscopy (XAS) and absorptive stripping voltammetry analyses, were used to determine the oxidation state and nature of aqueous Cr. A thermodynamic model, based on the Pitzer equations, was developed from the solubility measurements to account for dramatic increases in aqueous Cr with increases in NaOH concentration. The model includes only two aqueous Cr species, Cr(OH) 4 and Cr2O2(OH) 4 (although the possible presence of a small percentage of higher oligomers at >5.0 m NaOH cannot be discounted) and their ion–interaction parameters with Na+. The logarithms of the equilibrium constants for the reactions involving Cr(OH) 4 [Cr(OH)3(am) + OH Cr(OH) 4 ] and Cr2O2(OH) 4 2– [2Cr(OH)3(am) + 2OH Cr2O2(OH) 4 2– + 2H2O] were determined to be –4.36 ± 0.24 and –5.24 ± 0.24, respectively. This model was further tested and provided close agreement between the observed Cr concentrations in equilibrium with Cr(OH)3(am) in mixed NaOH–NaNO3 solutions and with high-level tank sludges leached with and primarily containing NaOH as the major electrolyte.  相似文献   

20.
Solubilities of arsenolite (As2O3, cub.) were measured from 22 to 90°C in water–acetone, water–acetic acid, and water—formic acid solutions of compositions ranging from the pure organic compound to pure water. Raman spectra were obtained at ambient temperature on As-containing water–acetic acid and water–acetone solutions. Results show that arsenic solvation by these organic compounds is negligible and hydroxide species dominate As speciation over a wide range of water activity (aH 2 O> 0.01). The solubility data were analyzed using an approach based on stoichiometric hydration reactions. Results show that As2O3 solubility can be described as a function of water activity, independently of the nature of the organic compound, by involving two neutral As hydroxide complexes: As(OH)3 and As(OH)3·4H2O. Stability constants derived for these species indicate that hydration weakens with increasing temperature. Calculations using these constants show that at low temperatures the tetrahydrate As(OH)3·4H2O is dominant in water-rich solutions; by contrast, in high-temperature crustal fluids, As(OH)3 becomes the major As species. The proposed hydration model can be used to analyze solubility of arsenic-bearing minerals and arsenic transport in complex H2O–CO2—electrolyte solutions encountered in natural and industrial environments.  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