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 203 毫秒
1.
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   

2.
The equilibria between tungsten(VI) and iminodiacetic acid (IDA) have been studied in aqueous solution. The stoichiometry and stability constants of the complexes formed are determined from a combination of potentiometric and Uv spectroscopic measurements. All measurements are carried out at 25°C, pH 7.5 and different ionic strengths ranging from (0.1 to 1.0) mol dm−3 (NaClO4). According to these results, tungsten(VI) forms a mononuclear complex with IDA of the type (WO3L2−). By introducing two empirical parameters C and D in the complex-formation reaction between tungsten(VI) and IDA, the dependence of the dissociation and stability constants on ionic strength is described by a modified Debye-Huckel-type equation. Finally, a pattern for the ionic strength dependence is obtained.  相似文献   

3.
The formation constants of species formed in the systems H+?+?W(VI)?+?nitrilotriacetic acid (NTA) and H+?+?NTA have been determined in aqueous solution for pH?=?4–9 at 25°C and different ionic strengths ranging from 0.1 to 1.0?mol?dm?3 NaClO4, using potentiometric and spectrophotometric techniques. It was shown that tungsten(VI) forms a mononuclear 1?:?1 complex with NTA of the type WO3L3? at pH?=?7.5. The composition of the complex was determined by the continuous variations method. The complexation of molybdenum(VI) with glutamic acid was investigated in aqueous solution ranging in pH from 4 to 9, using polarimetric, potentiometric and spectrophotometric techniques. The composition of the complex was determined by the continuous variations method. It was shown that molybdenum(VI) forms a mononuclear 1?:?1 complex with glutamic acid of the type MoO3L2? at pH?=?6.0. The dissociation constants of glutamic acid and the stability constants of the complex were determined at 25°C and at ionic strengths ranging from 0.1 to 1.0?mol?dm?3 sodium perchlorate. In both complex formation reactions the dependence of the dissociation and stability constants on ionic strength is described by a Debye-Huckel type equation. Finally, a comparison has been made between the patterns of ionic strength dependence for the two complexes and the results have been compared with data previously reported.  相似文献   

4.
The interaction between mellitate (benzene-1,2,3,4,5,6-hexacarboxylate) and diethylenetriamine and spermidine was studied by potentiometric and calorimetric techniques in aqueous solution at t = 25°C, moreover by diffractometric technique in the solid state. Thermodynamic parameters of several ApLHq species with p = 1, q = 2…5, for both systems, were calculated; in addition the following species were found: A2LH6, for A = dien; ALH6, for A = spd. The stabilities of these species are quite high; for example formation constants regarding the ALH33− species are: logK0 = 5.7 for A = dien and logK0 = 10.1 for A = spd. Thermodynamic parameters are generally function of charges involved in the formation reaction and some linear relationships were found. The speciation profiles show that the formation percentages, for the mononuclear species in particular, are high, in a wide pH range. The sequestering ability of mlt towards polyamines was analysed, also considering the already published data regarding ethylenediamine. By considering the formation constants at various ionic strength and the enthalpy values of triamine–mellitate species, it was possible to evaluate the dependence of sequestering ability on ionic strength, temperature and pH. Solid state investigation of two cocrystals, made by mellitate trianion and triammonium cations, provides information on the formation of supramolecular anionic layers. Comparison with the stability of mellitate–alkaline earth metals species was made.  相似文献   

5.
Kinetic characteristics of 19 transfer reactions of benzoyl group from N-benzoyloxypyridinium salts to pyridine N-oxides and 4-dimethylaminopyridine were studied in acetonitrile by the stopped-flow method. The rate of an identical reaction for 4-methoxypyridine was measured by dynamic NMR spectroscopy. For 5 other identical reactions the rates were estimated from Bronsted correlations. Equilibrium constants were estimated with the use of UV spectrophotometry (6), IR spectroscopy (2), from kinetic data (K ij = k ij /k ji ) (2), and in one case as logK i−j = logK i−x − logK j−x . The second order rate constants (k ij ) varied in the range 102–105 l mol−1 s−1, the equilibrium constants (K ij ) in the range 102–10−2; the activation parameters (ΔH ) were within 15–50 kJ mol−1, (−ΔS ) −20–110 J mol−1 K−1. The reactions under study occur in a single stage following the concerted SN2 mechanism through an early associative transition state. The benzoyl groups exchange rate and equilibrium are well described by simplified Marcus equation (omitting the quadratic term).__________Translated from Zhurnal Organicheskoi Khimii, Vol. 41, No. 5, 2005, pp. 788–792.Original Russian Text Copyright © 2005 by Rybachenko, Schroeder, Chotii, Lenska, Red’ko, Kovalenko.  相似文献   

6.
The Cu–sarcosine–OH and Cd–MIDA–OH systems have been studied by differential pulse polarography (DPP) at a fixed total ligand to total metal concentration ratio and varied pH at 298 K and μ=0.5 mol dm−3 in the background of NaNO3. Both the metal–ligand systems show initially dynamic (labile), followed by semi-dynamic behaviour on the DPP time scale. It has been shown that the experimental and calculated DPP complex formation curves used previously only for labile metal–ligand systems can be employed for the modelling of all species formed in a solution and optimisation of their stability constants. The stability constants of ML and ML2 complexes as log β were estimated for CuII and CdII as 7.75±0.02, 14.49±0.01 and 6.67 ±0.02, 12.00±0.02, respectively (all known hydroxide species of copper and cadmium, including polynuclear species, were incorporated into the metal–ligand–OH systems). The formation of the complex CuL2(OH) is suggested also and its stability constant as log β has been estimated to be 16.2±0.2. Results reported here seem to be reasonable when compared with the literature data reported at 298 K and different ionic strengths.  相似文献   

7.
Use of ascorbic acid as an indicator for the determination of the classical formation constants of uranyl association complexes with some ligand anions is studied in solutions of 0.05 ionic strength. The results enable the establishment of a sequence of complexation for the anions investigated vis., fluoride, chloride, bromide, and sulfate. This is as follows in order of decreasing complexing power on uranium: F > SO42− > Cl > Br. The validity of using the indirect ascorbate method for the determination of classical equilibrium constants is established for the uranyl sulfate complex by comparison of the K values with those obtained by the direct spectrophotometric procedure using the continuous variation method. Both procedures indicate the formation of a 1:1 uranyl sulfate complex.  相似文献   

8.
Complex formation in the Nb6O 19 8? -WO 4 2? -H+-H2O system with c Nb : c W = 1 : 5 and varied c Nb + W 0 = 10?2, 5 × 10?3, 2.5 × 10?3, and 10?3 mol/L) has been studied. Distribution diagrams were simulated for individual niobium(V) and tungsten(VI) isopolyanions and mixed isopolyniobotungstates for $Z = \frac{{c_{H^ + }^0 }}{{c_{Nb + W}^0 }} = 0 - 3.0$ in an NaCl background electrolyte. We have shown that isopolyniobotungstates-6 of composition H x NbW5O 19 (3 ? x)? are formed via H x Nb n W6?n O 19 (2 + n ? x)? (n=2, 3, 5) ions. The concentration formation constants and thermodynamic formation constants of isopolyniobotungstate anions (IPNTAs) in aqueous solution have been calculated. Salt Tl3NbW5O19·9H2O has been synthesized and identified by chemical analysis and IR spectroscopy.  相似文献   

9.
Positron lifetime measurements were performed on a series of organic electron acceptors and charge-transfer complexes in solution. The acceptors cause both positronium (Ps) inhibition (with maybe one exception) and quenching, but when an acceptor takes part in a charge-transfer complex the inhibition intensifies and the quenching almost vanishes. The reaction constants between ortho-Ps and the acceptors were determinded to be: 1.5 × 1010 M−1 s−1 for SO2 in dioxane 3.7 × 1010 M−1 s−1 for SO2 in n-heptane, 3.4 × 1010 M−1 s−1 for tetracyanoquinodimethane in tetrahydrofurane and 1.6 × 1010 M−1 s−1 for tetracyanoethylene in dioxane. From the ortho-Ps lifetimes in solutions containing charge-transfer complexes complexity constants were determined that were in reasonable agreement with constants obtained from optical data. The influence of acceptors and charge-transfers complexes on the Ps yield was interpreted in terms of the spur reaction model of Ps formation. Correlation was also made to gas phase reaction between electron acceptors and free electron, as well as to pulse radiolysis data.  相似文献   

10.
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   

11.
The formation of MTar, MCit2– and M2Cit complexes (M is Na+ or K+) was established in reactions of aqueous solutions of citric and tartaric acids with sodium or potassium chloride solutions; their stability constants were determined by potentiometric titration in aqueous solution at 298.15 K and ionic strength 0.1 and 0.3 mol/l with tetraethylammonium chloride (TEACl) as a supporting electrolyte. Heat effects of reactions between citric acid solutions and sodium or potassium chlorides were measured by calorimetric method at 288.15, 298.15, and 308.15 K and at the ionic strength 0.1, 0.2, and 0.3 M TEACl. The increasing ionic strength was found to decrease exothermic effect of complex formation processes, while the temperature produced the opposite effect. Extrapolation to zero ionic strength was used to find thermodynamic stability constants and standard heat effects of complex formation reactions in solutions of oxyacids. The changes in entropy and heat capacity were calculated as well as standard enthalpies of formation of Na and K complexes of the indicated oxyacids in aqueous solution at 298.15 K.__________Translated from Koordinatsionnaya Khimiya, Vol. 31, No. 4, 2005, pp. 253–261.Original Russian Text Copyright © 2005 by Zelenina, Zelenin.  相似文献   

12.
The formation of complexes between Mo(VI) and 8-hydroxy-quinoline (oxine) and four oxine derivatives were investigated by multiwavelength molecular absorption spectrometry, potentiometry, and polarography. The following pKOH- and pKNH- values of the ligands and logK 211-values of the complexes MoO2(OH)2L x (x=1 or 2) were obtained at 25° C and an ionic strength of 1M(NaClO4): 5,7-dinitro8-hydroxyquinoline 4.59, <0, 14.50; 7-nitro-8-hydroxyquinoline-5-sulfonic acid 5.34, 0.41, 15.70; 7-iodo-8-hydroxyquinoline-5-sulfonic acid 6.98, 2.62, 17.65; 8-hydroxyquinoline-5-sulfonic acid 8.33, 4.13, 18.71; and 8-hydroxyquinoline 9.62, 5.28, 19.69. A good linearity was found between logK 211 and the sum of the pK-values of the OH- and NH+-groups. The dependence of the peak current of Mo(VI)-determinations by adsorption polarography of the 7-nitro-8-hydroxyquinoline-5-sulfonate complex of Mo(VI) MoO(OH)3L can quantitatively be described at pH 0.8–2 using the corresponding pK-values and the log K311 of 18.54±0.03, determined by polarography.  相似文献   

13.
Bromophenol blue indicator was used in UV-visible spectrophotometric measurements to study ion association constants of alkali metal fluorides. The equilibrium constants for the ion pair formation of the alkali metal fluorides were determined as a function of ionic strength at one atmosphere pressure and 25°C. The effect of pressure on these association constants was measured at a constant total ionic strength of 1.0 mol-kg–1 over a pressure range of 1 to 2000 atmospheres at 25°C. The pressure dependences of the stoichiometric association constants of the alkali metal fluorides are given by: lnK LiF * =0.77–2.47×10–4P–2.12×10–8P2; lnK NaF * =0.53–1.08×10–4P–1.66×10–8P2; lnK KF * =0.24–4.41×10–5P–7.15×10–8P2; lnK RbF * =–0.17–8.65×10–5P–4.51×10–8P2; and lnK CsF * = –0.37–1.14×10–4P–6.82×10–8P2, where P is the pressure in atmospheres. The stoichiometric molar volume and compressibility changes for ion pair formation of the alkali metal fluorides were evaluated from the pressure dependence of K MF * data. The thermodynamic association constants were also calculated making use of activity coefficient data from the Pitzer equations. The partial molal volume and compressibility changes for ion pair formation of each alkali metal fluoride are reported.  相似文献   

14.
The 17O and 14N paramagnetic transverse relaxation time and chemical shift of proline as well as of water, in aqueous solutions of Co(II), Cu(II) and Mn(II) were measured as a function of pH, temperature, and metal ion concentration. The relaxation results were fitted to a theoretical equation linking the Swift-Connick equation to the stability constants of the major complexes in equilibrium. Stability constants for the major complexes of the three ions in this work were determined, along with thermodynamic parameters for some of the complexes. Two complexes of Co(II) were detected directly by 17O NMR at basic pH, and were assigned to CoPrO2 and CoPro3. The hyperfine coupling constant for these two complexes, A/h, was determined directly from the isotropic shift and was found to be −0.63 and −0.31 MHz, respectively. CoPrO2 could be detected in the pH range 6–12, for Co(II) concentrations greater than 0.04 M, and its chemical shift was around 700 ppm downfield from free proline, at 300 K. CoPro3 was detected only at pH 11, in the temperature range 275–284 K, with a chemical shift of 390 ppm downfield from free proline.  相似文献   

15.
The complex species formed in aqueous solution (25 C, I = 3.0 mol-dm−3 KCl ionic medium) between V3+ cation and the ligands: picolinic acid (Hpic, HL) and dipicolinic acid (H2dipic, H2L), have been studied potentiometrically and by spectrophotometric measurements. The application of the least-squares computer program LETAGROP to the experimental emf (H) data, taking into account the hydrolytic species of V3+ ion, indicates that under the employed experimental conditions, the formation of the complexes [VL]2+, [V(OH)L]+, [VL2]+, [VL3], [V2OL4] with picolinic acid and the complexes [VL]+, [V(OH)L], [V(OH)2L], [V(HL)(L)], and [VL2] with dipicolinic acid were observed. The stability constants of the complexes formed were determined by potentiometric measurements, and spectrophotometric measurements were done in order to perform a qualitative characterization of the complexes formed in aqueous solution.  相似文献   

16.
A detailed study of the electrochemical reduction of diacetylbenzene A in aqueous medium between Ho = −5 and pH 14 is presented. The reactants are strongly adsorbed, so that the reactions are of a surface nature. From Ho = −5 to pH 6, a global 2e reduction yielding an enediol-type intermediate occurs. Analysis using the theory of the square schemes with protonations at equilibrium shows that, up to pH 4, the reaction is controlled by the first electron uptake, the paths being successively H+e and eH+. The elementary electrochemical surface rate constants are 9.6 × 107 s and 1.2 × 106 s for AH+ and A respectively. From pH 6 to 14, a le adsorption wave, corresponding to the formation of (a) monoradical(s), appears and is followed by a le wave due to the reduction of the radical(s). A dimerization occurs, due to the coupling A + AH, as in the case of the monocarbonyl compounds. The rate of this surface process, kd = 5 × 1013 cm2 mol−1 s−1, is markedly smaller than the rate of the homogeneous reaction obtained in alkaline ethanol by Savéant et al. for the coupling of the radicals of benzaldehyde, benzophenone and acetophenone.  相似文献   

17.
18.
The ionic conductivity of solid solution Cd0.77Sr0.23F2 is 1.6 × 10−4 S/cm at 500 K. The conduction mechanism changes from a vacancy mechanism to an interstitial one at 523–553 K. In solid solutions Cd0.9R0.1F2.1 (R = La-Lu, Y), the activation enthalpy of conduction decreases from 0.9 to 0.8 eV with decreasing ionic radius of R3+, raising the 500-K conductivity from 6 ×10−6 S/cm for La3+to 6 × 10−5 S/cm for Lu3+. For crystalline Cd0.95In0.05F2.05, ionic and electronic conductivities at 313 K equal 5 × 10−4 and 5 − 10−6 S/cm.__________Translated from Elektrokhimiya, Vol. 41, No. 5, 2005, pp. 627–632.Original Russian Text Copyright © 2005 by Sorokin, Buchinskaya, Sul’yanova, Sobolev.  相似文献   

19.
The geometric structures of the ground state and metastable isomers of the nitroso complexes trans-[RuCl4(NO)L]q (L = H2O, NH3, pyrazine (Pz), q = −1; Cl, OH, CN, NO 2 , q = −2) and cis[RuCl4(NO)L]q (L = Pz, q = −1) were optimized in terms of the density functional theory. The variation of the trans-ligand L influences the relative energy of the metastable isomer with a side NO coordination. The presence of π-acceptor substituents in the trans-ligand L decreases the energy.__________Translated from Koordinatsionnaya Khimiya, Vol. 31, No. 8, 2005, pp. 606–611.Original Russian Text Copyright © 2005 by Lyubimova, Sizova.  相似文献   

20.
The solubility and the complex equilibria in the system Mn2+-CO2-H2O have been investigated at 25°C in solutions of the constant ionic strength 3 M (Na)ClO4. From experimental data the following values for equilibrium constants and Gibbs free energies of formation are deduced: A predominance area diagram for the system Mn2+-H2-CO2(g)-H2S(g) including MnCO3(ppt.), α-MnS and Mn2+ is given.  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