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 619 毫秒
1.
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   

2.
Solubility studies on UO2(c), precipitated at 90°C from low-pH U(IV) solutions, were conducted under rigidly controlled redox conditions maintained by EuCl2 as a function of pH and from the oversaturation direction. Samples were equilibrated for 24 days at 90°C and then for 1 day at 22°C. X-ray diffraction (XRD) analyses of the solid phases, along with the observed solubility behavior, identified UO2(c) as the dominant phase at pH1.2 and UO2(am) as the dominant phase at pH1.2. The UV-Vis-NIR spectra of the aqueous phases showed that aqueous uranium was present in the tetravalent state. Our ability to effectively maintain uranium in the tetravalent state during experiments and the recent availability of reliable values of Pitzer ion-interactionparameters for this system have helped to set reliable upper limits for the log K o value of –60.2 + 0.24 for the UO2(c) solubility [UO2(c) + 2H2O U4+ + 4OH] and of >–11.6 for the formation of U(OH)4(aq) [U4++ 4H2O U(OH)4(aq) + 4H+]  相似文献   

3.
The large number of possible chemical reactions represents a severe burdenfor modeling of even relatively simple plasma systems. Reduced sets ofchemical reactions have been obtained for numerical simulations of nitrogenand nitrogen-hydrogen plasma jets flowing into an atmospheric airenvironment. The important or active reactions are determined based on asimplified reduction method. A reaction is considered active if it leadsto higher sensitivities than a specified cutoff sensitivity of 1%. Theactive reactions exert a significant influence on main plasma parameters,such as velocity, temperature, and species concentrations. The sensitivityanalysis for the specified systems shows that two NO reactions, known asZel'dovich reactions (N2+ONO+N andNO+OO2+N),(1) are both active in a nitrogenplasma jet. On the other hand, the latter is not active and may be omittedin a nitrogen–hydrogen plasma jet. A nitrogen–hydrogen plasmajet requires contribution of two active charge exchange reactions:N2+N+N+ 2+N andN+H+N+ +H, while only the former is needed in a nitrogen plasmajet. The dissociation reactions are all active in both plasma jets, exceptthe dissociation of OH.  相似文献   

4.
Summary The kinetics of reversible complexation of NiII and CoII with iminodiacetato(pentaammine)cobalt(III), [(NH3)5-Co(idaH2)]3+ and NiII with iminodiacetato(tetraammine)-cobalt(III), [(NH3)4Co(idaH)]2+, have been investigated by the stopped-flow technique at 25 °C, pH = 5.7–6.9 and I = 0.3 mol dm –3. The reaction paths (NH3)5Co(idaH)2++M2+(NH3)5Co(ida)M3++H+ (NH3)5Co(ida)++M2+(NH3)5Co(ida)M3+ (NH3)4Co(ida)++Ni2+(NH3)4Co(ida)Ni3+ have been identified (idaH = N+H2(CH2CO2)2H, ida = NH(CH2COO)2–]. The rate parameters for the formation and dissociation of the binuclear species are reported. The data are essentially consistent with an I d mechanism. The dissociation rate constants of the binuclear species indicate that Ni2+ and Co2+ are chelated by the coordinated iminodiacetate moiety.  相似文献   

5.
Measurements of acoustic absorption and velocity as a function of frequency and concentration in KH2PO4–K2HPO4 buffers at 4°C and pH 5-7 are reported. The dependence of the observed acoustic relaxation parameters on concentration is consistent with that to be expected from perturbation of a monomer-dimer equilibrium with an equilibrium constant [for 2H2PO 4 (H2PO4)2 2–] of 0.21 M–1, a bimolecular rate constant of 5×108 M–1-sec–1 and a standard volume change of –5 cm3 mole. The equilibrium constant for H2PO 4 + HPO4 2–H3(PO4)2 3– is estimated to be 0.7 M–1.  相似文献   

6.
From several strontium distribution experiments with 85Sr tracer, the extraction constant corresponding to the equilibrium Ca2+(aq)+SrL2+(nb) CaL2+(nb)+Sr2+ (aq) in the two-phase water-nitrobenzene system (L = 18-crown-6; aq = aqueous phase, nb = nitrobenzene phase) was tentatively evaluated as log K ex (Ca2+,SrL2+) = –1.9±0.1. Furthermore, the stability constant of the calcium — 18-crown-6 complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log nb(Cal2+) = 10.1±0.1.  相似文献   

7.
Summary A study of the Cd(CN)2 +x X [Cd(CN)2X x ] x– equilibrium (where X = Cl, Br or CNS) has been carried out at 18° and 38° by measuring the solubility of cadmium cyanide in potassium chloride, bromide and thiocyanate at various concentrations, and at a high ionic strength (6 M) maintained with sodium perchlorate to minimise the effect of activity coefficients. Equilibrium constants forx = 1 and 2 have been calculated and clearly favour the situation wherex = 1. H values for the dissociation of [Cd(CN)2X] have also been calculated.  相似文献   

8.
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   

9.
The hydrolysis constants of telluric acid were determined by potentiometric titrations at 25°C andI=1.0 mol kg–1 NaClO4. Using these results the partial molar volume change according to the dissociation reaction Te(OH)6(aq) TeO(OH) 5(aq) +H (aq) + was measured densitymetrically.
Das Dissoziationsvolumen der Tellursäure (Kurze Mitteilung)
Zusammenfassung Die Hydrolysekonstanten der Tellursäure wurden bei 25°C undI=1.0 mol kg–1 NaClO4 durch potentiometrische Titrationen bestimmt. Diese Ergebnisse wurden verwendet, um die Volumsänderung zufolge der Dissoziationsreaktion Te(OH)6(aq) TeO(OH) 5(aq) +H (aq) + durch Dichtemessungen zu ermitteln.
  相似文献   

10.
PuO2(am) solubility was investigated as a function of time, for pH from 0.5 to 11, and in the presence of 0.001 M FeCl2 or 0.00052 M hydroquinone to determine the effect of environmentally important reducing agents on PuO2(am) solubilization under geological conditions. Equilibrium was reached in <4 days. The observed PuO2(am) solubilities were many orders of magnitude higher than the Pu(IV) concentrations predicted from thermodynamic data. Spectroscopic, solvent extraction, and thermodynamic analyses of data showed that Pu(III) was the dominant aqueous oxidation state. The experimental pH, pe, and Pu(III) concentrations from both the Fe(II) and hydroquinone systems provided a log K 0 value of 15.5 ± 0.7 for [PuO2(am) + 4H+ + e Pu3+ + 2H2O]. The data show that reduction reactions involving Fe(II) and hydroquinone are relatively rapid and that reductive dissolution of PuO2(am), hitherto ignored, may play an important role in controlling Pu behavior under reducing environmental conditions.  相似文献   

11.
The precipitation and dissolution of AgI in the presence of methylene blue (MB · Cl) and sodium dodecyl sulphate (NaDS) was followed by X-ray diffraction analysis.At high MB · Cl concentration, the absence of AgI precipitates was observed, which is explained by considering the redox process MB+ + 2I MB + I2. The decrease in I concentration causes dissolution or inhibition of growth of solid AgI which is significant at relatively high MB · Cl concentrations. The addition of NaDS causes the disappearance of these effects, which is explained by the incorporation of MB in NaDS micelles.These explanations are supported by the potentiometric measurements using Ag/Ag2S and Pt electrodes.  相似文献   

12.
The solubility of Na2SO4 (s) (thenardite) and the interactions between magnetiteand aqueous Na2SO4 near the critical point of water have been determined in azirconium-alloy flow reactor at temperatures 350°C t 375°C and isobaricpressures 190 p 305 bar. The experimental solubility data are describedwell as a function of temperature and solvent density 1 byln x(Na2SO4, aq.) = –10.47 – 27550/T +(4805/T) ln 1.The interaction between magnetite and Na2SO4 (aq.) was examined from 250 to370°C at molalities near the saturation composition of Na2SO4 (s). While no solidreaction products were observed, HS (aq.) was observed to form above 350°Cby sulfate reduction, as a product of the reaction8 Fe3O4(s) + Na2SO4 (aq.) + H2O(l)= 12 Fe2O3 (s) + NaHS (aq.) + NaOH (aq.).The reduction reaction appears to be controlled by surface reaction kinetics, ata level well below the equilibrium molality of HS (aq.). Metallic iron reactedwith Na2SO4 (aq.) in a similar fashion at temperatures above 350°C, to yieldhigher molalities of HS (aq.).  相似文献   

13.
Zusammenfassung Die Photolyse des Wassers bei 1849 Å wurde unter Verwendung von 0,01m-Formiat als Fänger für die H-Atome und OH-Radikale untersucht. Dabei diente 5m-Äthanol als Aktinometer mit einem korrigierten Wert für (H2)=0,50. In diesem Fall wurde eine Quantenausbeute der Wasserphotolyse (H, OH)=0,36±0,01 bestimmt. Bezieht man die exper. Daten auf das N2O-Aktinometer bei (–N2O)=1,0, dann ist (H, OH)=0,29±0,01. In diesem Wert ist auch die Quantenausbeute der reaktionsfähigen angeregten Wassermoleküle, die mit Formiat reagieren, inbegriffen. Auf Grund von experimentellen Daten wurde ferner die Bildung von solvatisierten Elektronen (e aq) vorgeschlagen. Durch Sättigung der Formiatlösung mit Kohlensäure, die sowohl vone aq als auch von H2O* reduziert werden kann, wurde (e aq, H2O*)>0,02<0,04 bestimmt.
Liquid water photolysis at 1849 Å was investigated by using 0,01m-formate as scavenger for the H and OH radicals. 5M-ethyl alcohol serviced as actinometer with a corrected value of (H2)=0,50. The quantum yield of water photolysis was determined in this case to be (H, OH)=0,36±0,01. When the experimental results are related to N2O actinometer with (–N2O)=1,0, a quantum yield of (H, OH)=0,29±0,01 is obtained. This value includes also the quantum yield of the excited water molecules which react with the formate. Based on experimental data the formation of solvated electrons (e aq) is proposed. By saturation of the formate solution with carbon dioxide, which can be reduced bye aq as well as by H2O*, (e aq, H2O*>0,02<0,04 was determined.


Mit 4 Abbildungen  相似文献   

14.
Summary Reaction of 5,7-dioxo-1,4,8,11-tetra-azacyclotetradecane with acrylonitrile gives the dicyanoethylated ligand (L). The CuII complex [CuLH-2]·2H2O has been isolated from basic solution where the macrocycle is deprotonated and acts as a dinegative quadridentate ligand. The ligand L is protonated in acidic solution and the ionisation equilibria can be summarised as LH inf2 sup2+ LH+ +H+; K1 LH+ L + H+; K2 where pK1 = 3.05 and pK2 = 5.94 at 25 °C and I = 0.1 mol dm-3 (NaNO3). Complexation with CuII can be represented by the equilibria at 25 °C. Cu2+ + L [CuLH-1]+ + H+; log11 – 1 = -3.43 Cu2+ + L [CuLH-2] + 2H2+; log11 – 2 = -9.18 For NiII only the single equilibrium is of importance. Ni2+ + L [NiLH-2] + 2H2+; log11 – 2 = -14.45  相似文献   

15.
Fe(III) hydrolysis and fluoride complexation behavior was examined in 0.68 molal sodium perchlorate at 25°C. Our assessment of the complexation of Fe(III) by fluoride ions produced the following results: logF1 = 5.155, logF2 = 9.107, logF3 = 11.96, logF4 = 13.75, where logFn = 5.155=[FeF n (3-n)+ ][Fe3+]–1[F]–n. The stepwise fluoride complexation constants,FK n+1, obtained in our work (where logF K n+1 =logFn) indicate that K n+1/K n =0.072±0.01. Formation constants for equilibria, Fe3++nH2OFe(OH) n (3–n)+ +nH+, expressed in the form n * [Fe(OH) n (3-n)+ ][H+]n ,[Fe 3+]-1, were estimated as 1 * = –2.754, and 2 * –7. Our study indicates that the results of previous hydrolysis investigations include very large overestimates of Fe(OH) 2 + formation constants.  相似文献   

16.
The binary system H2O—UO2(NO3)2 was studied by solubility measurements and constant heat flow thermal analysis. Temperature and composition of the eutectic transformation between ice and uranyl nitrate hexahydrate were accurately defined. A new hydrate with 24 molecules of water decomposes at –21°C according to the peritectoid reaction<UO2(NO3)2·24H2O> <UO2(NO3)2·6H2O> + 18<H2O>The quasi-ideal model was applied to the solid—liquid equilibria, using the following reaction hypothesis:((UO 2 2+ )) + 2((NO 3 ))+ h((H2O)) ((UO2OH+aq)) + ((H3O+aq + 2((NO 3 aq))A complete calculation of the binary system was carried out with a global ionic hydration number h equal to 9 in the aqueous solutions. It allowed to the melting enthalpies of uranyl nitrate hydrates.
This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

17.
Using flow microcalorimetry, the ion association reaction M2+(aq)+Fe(CN) 6 4– (aq)=MFe(CN) 6 2– (aq) (M=Ca, Mg) has been studied at 25°C over the ionic strength range 0.02 to 0.08 mol-dm–3. Analyses of the data to obtain Ho, the enthalpy change at infinite dilution, are described. The value obtained for Ho is sensitive to the kind of functions used to correct for non-ideal behavior.  相似文献   

18.
An ion-exchange method was used to determine complexation constants for the Ni-oxalate and Ni-carbonate systems in a NaClO4 background electrolyte. The Ni-oxalate data were interpreted in terms of a single Niox(aq) complex having log K 1 values for Ni2+ + ox2– Niox(aq) of 3.9 ± 0.1 (I.S. = 0.5 mol-L–1 p[H] = 7.1) and 4.4 ± 0.1 (I.S. = 0.1 mol-L–1 p[H] = 8.6) at 22 ± 1C. Specific ion-interaction theory (SIT) was used to obtain log K 1 = 5.17 ± 0.05 (95% confidence level and = –0.23 ± 0.15) at I.S. = 0. The Ni-carbonate studies were carried out at p[H] values of 7.5, 8.5, and 9.6 in 0.5 mol-L–1 NaClO4/NaHCO3 solutions. The NiCO3(aq) species was the dominant complex in the [CO3 2–] concentration ranges studied at all three p[H] values. A log K 1 value for Ni2+ + CO3 2– NiCO3(aq) of 2.9 ± 0.3 was deduced at I.S. = 0.5 mol-L–1. Extrapolating this value to zero ionic strength using the SIT approach yielded log K 1 = 4.2 ± 0.3 (95% confidence level and = –0.26 ± 0.04). The data allowed upper bound values for the complexation constants for NiHCO3 + and Ni(CO3)2 2– to be estimated, i.e., log K < 1.4 for Ni2+ + HCO3 NiHCO3 +, and log K 2 < 2 for NiCO3(aq) + CO3 2– Ni(CO3)2 2–, respectively.  相似文献   

19.
The influence of elevated temperatures on the formation of 1:1 chloro complexes for Eu3+ and Am3+ are reported. Using a solvent extraction technique, stability constants for the equilibrium M(aq) 3++Cl(aq) MCl(aq) 2+ have been measured in the temperature range of 25–75 °C. Modest increases in 1 are observed, and small positive enthalpies for these reactions are reported. These data are discussed in the context of previous reports for the trivalent lanthanide and actinide chloro systems.  相似文献   

20.
Prior to this study there were no thermodynamic data for isosaccharinate (ISA) complexes of Fe(III) in the environmental range of pH (>~4.5). This study was undertaken to obtain such data in order to predict Fe(III) behavior in the presence of ISA. The solubility of Fe(OH)3(2-line ferrihydrite), referred to as Fe(OH)3(s), was studied at 22?±?2?°C in: (1) very acidic (0.01?mol·dm?3 H+) to highly alkaline conditions (3?mol·dm?3 NaOH) as a function of time (11?C421?days), and fixed concentrations of 0.01 or 0.001?mol·dm?3 NaISA; and (2) as a function of NaISA concentrations ranging from approximately 0.0001 to 0.256?mol·dm?3 and at fixed pH values of approximately 4.5 and 11.6 to determine the ISA complexes of Fe(III). The data were interpreted using the SIT model that included previously reported stability constants for $ {{\text{Fe(ISA}})_{n}}^{3 - n} $ (with n varying from 1 to 4) and Fe(III)?COH complexes, and the solubility product for Fe(OH)3(s) along with the values for two additional complexes (Fe(OH)2(ISA)(aq) and $ {\text{Fe(OH)}}_{ 3} ( {{\text{ISA}})_{2}}^{2 - } $ ) determined in this study. These extensive data provided a log10 K 0 value of 1.55?±?0.38 for the reaction $ ({\text{Fe}}^{ 3+ } + {\text{ISA}}^{-} + 2 {\text{H}}_{ 2} {\text{O}} \rightleftarrows {\text{Fe(OH}})_{ 2} {\text{ISA(aq}}) + 2 {\text{H}}^{ + } ) $ and a value of ?3.27?±?0.32 for the reaction $ ({\text{Fe}}^{ 3+ } + 2 {\text{ISA}}^{-} + 3 {\text{H}}_{ 2} {\text{O}} \rightleftarrows {\text{Fe(OH)}}_{ 3} ( {\text{ISA}})_{2}^{2 - } + 3 {\text{H}}^{ + } ) $ and show that ISA forms strong complexes with Fe(III) which significantly increase the Fe(OH)3(s) solubility at pH?<~12. Thermodynamic calculations show that competition of Fe(III) with tetravalent ions for ISA does not significantly affect the solubilities of tetravalent hydrous oxides (e.g., Th and Np(IV)) in ISA solutions.  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