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 760 毫秒
1.
The relative sound speed of mixtures of aqueous solutions of NaCl–MgSO4 and MgCl2–Na2SO4 at I=0.1 and 0.5m have been determined at 5, 15, and 25°C and pressures to 1000 bars. The resulting sound speeds, adiabatic and apparent molal compressibilities have been compared to results estimated from binary solutions using an additivity principle — Young's rule. The estimated sound speeds agree with the measured values for the NaCl–MgSO4 system to ±0.15 m-sec–1 and for the Na2SO4–MgCL2 system to ±0.20 m-sec–1. The deviations increase with increasing ionic strength (±0.08 m-sec–1 at I=0.1 and ±0.25 m-sec–1 at I=0.5 m).The sound speed of seawater have also been estimated from 0 to 40°C, 0.1 to 0.7 ionic strength and 0 to 1000 bars. The estimates were found to be in good agreement (±0.4 m-sec–1) with the measured values.These results indicate that reasonable estimates of the adiabatic PVT properties of dilute mixtures of electrolyte solutions can be made using the additivity principle, without excess mixing terms.  相似文献   

2.
The speed of sound of mixtures of the six possible combinations of the major sea salt ions (Na+, Mg2+, Cl, and SO 4 2– ) have been determined at I=3.0 and at 25°C. The results have been used to determine the changes in the adiabatic compressibility of mixing Km the major sea salts. The values of Km have been fit to the equation Km=y2y3I2[k0+k1(1-2y3)] where yi is the ionic strength fraction of solute i, k0 and k1 are parameters related to the interactions of like-charged ions. The Young cross-square rule is obeyed to within ±0.04×10–6 cm3-kg–1-bar–1. A linear correlation was found between the compressibility k0 and volume v0 interaction parameters (104k0=–0.24+3.999 v0, s=0.15) in agreement with out earlier findings. Estimates of the sound speeds for the cross square mixtures (NaCl+MgSO4 and MgCl2+Na2SO4) were made using the equations of Reilly and Wood. The estimated sound speeds were found to agree on the average with the measured values to ±0.36 m-sec–1.  相似文献   

3.
The relative sound velocities (U-U°) of aqueous NaCl, Na2SO4, MgCl2, and MgSO4 solutions were measured from 0.05m to saturation and from 0 to 45°C. The sound speeds were combined with our earlier work and fitted to a function of molality and temperature to standard deviations within 0.3 m-sec–1. The adiabatic compressibilities, s, were determined from the sound speeds and used to calculate adiabatic apparent molar compressibilities, K,s, isothermal compressibilities, , and apparent molar compressibilities, K, were determined from the adiabatic values using literature data for expansibilities and heat capacities. The values of K have been extrapolated to infinite dilution using an extended limiting law. The resulting K0 at various temperatures are in reasonable agreement with literature values. The results of this study have been combined with our earlier results to derive a secant bulk modulus equation of state for NaCl, Na2SO4, MgCl2, and MgSO4 solutions valid from 0 to 50°C and 0 to 1000 bar.  相似文献   

4.
Heats of dilution of concentrated aqueous solutions (4.43 moles-kg–1) of FeCl2 were measured at 15, 25, and 35°C. The heat capacities of these concentrated solutions were also measured at the same temperatures. From these data the partial molal heat capacity, C p2 0 (FeCl2, aq, 298.15°K)=–2.56±30 J–°K–1–mole–1, was calculated. The partial molal heat capacity of Fe2+(aq), –2±30 J-°K–1-mole–1, was correlated with the correspondence principle equations of Criss and Cobble.  相似文献   

5.
The thermal behaviour of Ba[Cu(C2O4)2(H2O)]·5H2O in N2 and in O2 has been examined using thermogravimetry (TG) and differential scanning calorimetry (DSC). The dehydration starts at relatively low temperatures (about 80°C), but continues until the onset of the decomposition (about 280°C). The decomposition takes place in two major stages (onsets 280 and 390°C). The mass of the intermediate after the first stage corresponded to the formation of barium oxalate and copper metal and, after the second stage, to the formation of barium carbonate and copper metal. The enthalpy for the dehydration was found to be 311±30 kJ mol–1 (or 52±5 kJ (mol of H2O)–1). The overall enthalpy change for the decomposition of Ba[Cu(C2O4)2] in N2 was estimated from the combined area of the peaks of the DSC curve as –347 kJ mol–1. The kinetics of the thermal dehydration and decomposition were studied using isothermal TG. The dehydration was strongly deceleratory and the -time curves could be described by the three dimensional diffusion (D3) model. The values of the activation energy and the pre-exponential factor for the dehydration were 125±4 kJ mol–1 and (1.38±0.08)×1015 min–1, respectively. The decomposition was complex, consisting of at least two concurrent processes. The decomposition was analysed in terms of two overlapping deceleratory processes. One process was fast and could be described by the contracting-geometry model withn=5. The other process was slow and could also be described by the contracting-geometry model, but withn=2.The values ofE a andA were 206±23 kJ mol–1 and (2.2±0.5)×1019 min–1, respectively, for the fast process, and 259±37 kJ mol–1 and (6.3±1.8)×1023 min–1, respectively, for the slow process.Dedicated to Prof. Menachem Steinberg on the occasion of his 65th birthday  相似文献   

6.
The intradiffusion coefficients of Na+, Cl ions and water and the tracerdiffusion coefficients of Ca2+ ion have been measured in the ternary system NaCl–MgCl2–H2O at 25°C. The intradiffusion coefficients of Mg2+ in this system have been estimated from the corresponding Ca2+ diffusion measurements. Viscosities were measured at the same solution concentrations as were used for the diffusion experiments. Intradiffusion and tracerdiffusion coefficients in a range of temperatures from 5 to 45°C are reported for standard sea-water which is a member of the above ternary set.  相似文献   

7.
Apparent molar volumes and heat capacities of aqueous GaCl3 have been measured at 25°C in binary GaCl3 solutions up to 3 mol-kg–1, and in ternary GaCl3-HCl solutions, containing 0.1345 mol-kg–1 HCl to suppress hydrolysis, up to a concentration of 1 mol-kg–1 GaCl3. Using the Pitzer interaction model for the excess properties, and using ridge regression for the derivation of physically meaningful regression parameters, the measurements yield the following results for the standard molar properties and Pitzer parameters at 25°C: V0(GaCl3)=12.85 cm3-mol–1; 0 v (GaCl3)=1.10×10–4 kg-mol–1–J–1–cm–3; v 1 (GaCl3)=2.12×10–3 kg–mol–1–J–1–cm3; Cv(GaCl3)=1.34×10–5 kg2–J–1–cm3; Vo(GaOHCl2)=13.84 cm3–mol–1; C o p (GaCl3)=–480.8 J–K–1–mol–1; J 0 (GaCl3)=–8.02×10–6 kg–mol–1–K–2; J 1 (GaCl3)=0.73×10–4 kg–mol–1–K–2; CJ(GaCl3)=–2.52×10–6 kg2-mol–2-K–2; C p 0 (GaOHCl2)=20.4 J-K–1-mol–1. The latter parameter has only mathematical significance, its physical meaning is unclear. Comparison of the present experimental results for the standard molar properties of Ga3+ with semi-empirical correlations casts doubt upon the general validity of these correlation methods for trivalent cations.  相似文献   

8.
Equations in the ion-interaction (Pitzer) system are derived for the volume change on mixing any combination of the sea salts NaCl, Na2SO4, MgSO4, MgCl2 at constant ionic strenth. For these mixings of different charge types, the equations include complex differences of pure electrolyte terms. Recently measured data for each of the pure electrolytes provide these pure electrolyte terms. Other recent measurements on the volume change on mixing are compared with values calculated from the equations. At 25°C there is no need to introduce the mixing terms based on differences in the interactions of ions of the same sign. At other temperatures, the agreement without the mixing terms is good, but significant improvement is obtained by inclusion of the binary mixing terms Cl,SO 4 v and Na,Mg v . The equations and parameters can then predict the volumetric properties of any mixed solution of these salts over the range 0–100°C and to at least 3 mol-kg–1 ionic strength.  相似文献   

9.
A nanoceramic product of the composition Lu2Ti2O7 is synthesized by a coprecipitation method with a subsequent sublimation drying and an annealing at 650–1650°C. The conduction of Lu2Ti2O7 synthesized at 1650°C is ionic (10–3 S cm–1 at 800°C). Thus, a new material with a high ionic conduction has been discovered. The ordering in Lu2Ti2O7 is studied by methods of RFA, RSA, IK spectroscopy, electron microscopy, and impedance spectroscopy. The existence of a low-temperature phase transition fluorite-pyrochlore at 800°C and a high-temperature conversion order-disorder at 1650°C are established.Translated from Elektrokhimiya, Vol. 41, No. 3, 2005, pp. 298–303.Original Russian Text Copyright © 2005 by Shlyakhtina, Ukshe, Shcherbakova.  相似文献   

10.
The densities of mixtures of the six possible combinatons of the major sea salts (NaCl, Na2SO4, MgSO4, and MgCl2) were determined at constant ionic strengths of I=1.0 and I=3.0 at 25°C. The results are used to determine the volume changes for mixing (V m ) the major sea salts. The values of V m were fit to equations of the form V m where y i is the molal ionic strength fraction of solute i, and 0 and 1 are parameters related to the interaction of like-charged ions. The cross-square rule was found to hold at both ionic strengths. Density estimates were made without and with the addition of volume of mixing terms to Young's Rule and compared to the experimental values. The densities calculated with the addition of volume of mixing terms gave better estimates, demonstrating that the densities of concentrated brines can be more accurately estimated using V m terms. The equations of Reilly and Wood which include the cross-square rule were used to estimate the densities of the cross mixtures (NaCl–MgSO4 and MgCl2–Na2SO4). The estimated densities agree with the measured values to within ±30 ppm at I=1.0 and ±125 ppm at I=3.0.  相似文献   

11.
The apparent molar heat capacities, Cp,, of alkaline aqueous solutions of aluminum ion in excess NaOH have been measured at temperatures between 50 and 250°C in the overall molality range 0.3–1.7 mol-kg–1. Enthalpies of dilution, L, have also been determined at 99°C and apparent molar relative enthalpies, L, were calculated starting from 2.16 mol-kg–1 as the maximum concentration. Measurements of the above quantities have been performed by means of a differential flow calorimeter built in our laboratory and already described. The thermodynamic data obtained and the corresponding quantities for aqueous NaOH previously determined have been fitted to the equations of the Pitzer ionic interaction model to obtain parameters relative to aqueous NaAl(OH)4. These parameters permit the calculation of Cp,, and L for this species over the examined range of temperatures and concentrations.  相似文献   

12.
Electrochemistry of a mixture of hydrofullerenes C70H36—46 composed of C70H36, C70H38, C70H44, and C70H46 (50, 20, 14, and 15%, respectively) was studied by cyclic voltammetry in THF and CH2Cl2 in the –43—–13 °C temperature range. Two cathodic peaks, namely, one-electron reversible (E° = –3.16 V (Fc0/+), Fc is ferrocene) and irreversible (E p = –3.37 V (Fc0/+)) were observed for this mixture in THF. The irreversible broad oxidation peak (E p = 1.22 V (Fc0/+)) was observed in CH2Cl2. The reversible reduction peak (E° = –3.16 V) and irreversible oxidation peak (E p = 1.22 V) were attributed to the most stable hydrofullerene C70H36. The irreversible reduction (E p = –3.37 V) and oxidation (E p = 1.22 V) peaks were attributed to hydrofullerenes C70H44—46 with a higher degree of hydrogenation. The values of an electrochemical gap, which is an analog of the energy gap (HOMO—LUMO), are 4.38 and 4.59 V for C70H36 and C70H44—46, respectively, and indicate that these hydrofullerenes are sufficiently hard molecules with low reactivity in redox reactions.  相似文献   

13.
Apparent molar heat capacities and volumes have been determined for aqueous solutions of the mixed electrolytes Na5DTPA + NaOH, Na3CuDTPA + NaOH, and NaCu2DTPA + NaOH, and the single electrolyte Na3H2DTPA (DTPA=diethylenetriaminepentaacetic acid) at temperatures from 10 to 55°C. The experimental results have been analyzed in terms of Young's rule with the Guggenheim form of the extended Debye–Hückel equation and the Pitzer ion-interaction model. These calculations led to standard partial molar heat capacities and volumes for the species H2DTPA3–(aq), DTPA5–(aq), CuDTPA3–(aq), and Cu2DTPA(aq) at each temperature. The partial molar properties at 0.1 m ionic strength were also calculated. The standard partial molar properties were extrapolated to elevated temperatures with the revised Helgeson–Kirkham–Flowers (HKF) model. Values for the partial molar heat capacities from the HKF model have been combined with the literature data to estimate the ionization constants of H2DTPA3–(aq) and the formation constant of the CuDTPA3–(aq) copper complex at temperatures up to 300°C.  相似文献   

14.
The solubility of CaSO3·1/2H2O(c) was studied under alkaline conditions (pH>8.2), in deaerated and deoxygenated Na2SO3 solutions ranging in concentration from 0.0002 to 0.4M and in CaCl2 solutions ranging in concentration from 0.0002 to 0.01M, for equilibration periods ranging from 1 to 7 days. Equilibrium was approached from both the over- and the under-saturation directions. In all cases, equilibrium was reached in <1 days. The aqueous Ca2+–SO 3 2– ion interactions can be satisfactorily modeled using either ion-association or ion-interaction aqueous thermodynamic models. In the ion-association model, the log K°=2.62±0.07 for Ca2++SO 3 2– CaSO 3 0 . In the Pitzer ion-interaction model, the binary parameters (0) and (1) for Ca2+–SO 4 2– were used, and the value of (2) was determined from the experimental data. As expected given the strong association constant, the value of (0) was quite small (about –134). We feel a combination of the two models is most useful. The logarithm of the thermodynamic equilibrium constant (K°) of the CaSO3·1/2H2O(c) solubility reaction (CaSO3·1/2H2O(c)Ca2++SO 3 2+ +0.5H2O) was found to be –6.64±0.07.  相似文献   

15.
The composition of the products of carbothermal reduction of Sc2O3 is examined by X-ray diffraction and chemical analysis and by the hydrolysis method. At pressures of 10–2-1 Pa, the reaction starts in the temperature region of 1 000–1 200°C. The first product is Sc2OC of NaCl type; at 1 Pa and 1 400–1 500°C this substance is formed quantitatively (according to stoichiometry) within 50–100 h, repeated homogenization, however, is necessary, or else Sc2OC reacts locally with Sc2O3 giving Sc2O1+x C1–x . The lattice parameter of Sc2OC in the presence of Sc15C19 is 457.63pm. At temperatures above 1 500°C, Sc15C19 is incompletely formed by subsequent reaction with carbon. The product melts at cca. 1 800°C; carbon dissolves and the final composition approaches ScC2. Carbon separates during solidification. The phase fractions in the products are affected by evaporation, the vapour pressures above both Sc2OC and Sc15C19 being comparable with the pressure requisite for the carboreduction process. The results are discussed with respect to the often ambiguous published data.
Untersuchung hydrolysierbarer Carbide, XXII. Die karbothermische Reduktion von Scandiumoxid
Zusammenfassung Zur Untersuchung von Produkten der karbothermischen Reduktion von Scandiumoxid wurden sowohl röntgenographische und chemische Analyse als auch die hydrolytische Methode verwendet. Bei 10–2 bis 1 Pa beginnt die Reaktion im Temperaturintervall von 1 000–1 200°C. Das erste Reaktionsprodukt ist das Scandiumoxidcarbid Sc2OC von NaCl-Typ, das quantitativ (der Stöchiometrie entsprechend) bei 1 400–1 500°C und 1 Pa in 50–100 Stunden entsteht. Eine wiederholte Homogenisierung ist allerdings notwendig, damit es zu keiner lokalen Reaktion zwischen Sc2OC und Sc2O3 kommt, bei der dann die Phase Sc2O1+x C1–x entsteht. In Gegenwart von Sc15C19 ist der Gitterparameter von Sc2OCa=457.63pm. Über 1 500°C führt eine weitere Reaktion mit Kohlenstoff zu einer nicht ganz vollendeten Bildung von Sc15C19. Bei cca. 1 800°C schmilzt das Reaktionsprodukt bei gleichzeitiger Auflösung von weiterem Kohlenstoff und die Zusammensetzung nähert sich der Formel ScC2, beim Erstarren fällt der Kohlenstoff wieder aus. Die Verteilung der Phasen im Produkt wird von der Verdampfung beeinflußt, da die Dampfdrucke von Sc2OC und Sc15C19 mit dem zum Karboreduktionverlauf notwendigen Druck vergleichbar sind. Die erhaltenen Ergebnisse werden in Relation mit den nicht eindeutigen Angaben in der Literatur diskutiert.
  相似文献   

16.
Raman and IR data for aqueous CdSO4 and (NH4)2SO4 solutions have been recorded over broad concentration and temperature ranges. Whereas the v1-SO 4 2– band profile is symmetrical in (NH4)2SO4 solutions, in CdSO4 solutions a shoulder appears on the high frequency side which increases in intensity with increasing concentration and temperature. The molar scattering coefficient of the v1-SO 4 2– band is the same for all forms of sulfate in (NH4)2SO4 and CdSO4 solutions and is independent of temperature up to 99°C. The high frequency shoulder is attributed to the formation of a contact ion pair [Cd2+OSO 3 2– ] (11 associate). Also the v3-SO 4 2– antisymmetric stretching mode shows a splitting in the CdSO4 solution. Further spectroscopic evidence for contact ion pair formation is provided by IR spectroscopy. No higher associates or anionic complexes are required to interpret the spectroscopic data. The degree of association has been measured as a function of concentration and temperature. The thermodynamic association constant, KA=0.15±0.05 kg-mol–1 at 25°C is estimated from the Raman data by an extrapolation procedure by taking account of the activity coefficients. Values are reported for the activity coefficient of the ion pair. From the Raman temperature dependence studies, the enthalpy of formation for the contact ion pair is estimated to be 10±1 kJ-mol–1.  相似文献   

17.
The first and second molal dissociation quotients of succinic acid were measured potentiometrically with a hydrogen-electrode, concentration cell. These measurements were carried out from 0 to 225°C over 25° intervals at five ionic strengths ranging from 0.1 to 5.0 molal (NaCl). The dissociation quotients from this and two other studies were combined and treated with empirical equations to yield the following thermodynamic quantities for the first acid dissociation equilibrium at 25°C: log K1a=–4.210±0.003; H 1a 0 =2.9±0.2 kJ-mol–1; S 1a 0 =–71±1 J-mol–1-K–1; and C p1a 0 =–98±3 J-mol–1-K–1; and for the second acid dissociation equilibrium at 25°C: log K2a=–5.638±0.001; H 2a 0 = –0.5±0.1 kJ-mol–1; S 2a 0 =–109.7±0.4 J-mol–1-K–1; and C p2a 0 = –215±8 J-mol–1-K–1.  相似文献   

18.
Raman spectra of aqueous Zn(II)–perchlorate solutions were measured over broad concentration (0.50–3.54 mol-L–1) and temperature (25–120°C) ranges. The weak polarized band at 390 cm–1 and two depolarized modes at 270 and 214 cm–1 have been assigned to 1(a 1g), 2(e g), and 5(f 2g) of the zinc–hexaaqua ion. The infrared-active mode at 365 cm–1 has been assigned to 3(f 1u). The vibrational analysis of the species [Zn(OH2) 2 + ] was done on the basis of O h symmetry (OH2 as point mass). The polarized mode 1(a 1g)-ZnO6 has been followed over the full temperature range and band parameters (band maximum, full width at half height, and intensity) have been examined. The position of the 1(a 1g)-ZnO6 mode shifts only about 4 cm–1 to lower frequencies and broadens by about 32 cm–1 for a 95°C temperature increase. The Raman spectroscopic data suggest that the hexaaqua–Zn(II) ion is thermodynamically stable in perchlorate solution over the temperature and concentration range measured. These findings are in contrast to ZnSO4 solutions, recently measured by one of us, where sulfate replaces a water molecule of the first hydration sphere. Ab initio geometry optimizations and frequency calculations of [Zn(OH2) 2 + ] were carried out at the Hartree–Fock and second-order Møller–Plesset levels of theory, using various basis sets up to 6-31 + G*. The global minimum structure of the hexaaqua–Zn(II) species corresponds with symmetry T h. The unscaled vibrational frequencies of the [Zn(OH2) 2 + ] are reported. The unscaled vibrational frequencies of the ZnO6, unit are lower than the experimental frequencies (ca. 15%), but scaling the frequencies reproduces the measured frequencies. The theoretical binding enthalpy for [Zn(OH2) 2 + ] was calculated and accounts for ca. 66% of the experimental single-ion hydration enthalpy for Zn(II).Ab initio geometry optimizations and frequency calculations are also reported for a [Zn(OH2) 2 18 ] (Zn[6 + 12]) cluster with 6 water molecules in the first sphere and 12 in the second sphere. The global minimum corresponds with T symmetry. Calculated frequencies of the zinc [6 + 12] cluster correspond well with the observed frequencies in solution. The 1-ZnO6 (unscaled) mode occurs at 388 cm–1 almost in perfect correspondence to the experimental value. The theoretical binding enthalpy for [Zn(OH2) 2 18 ] was calculated and is very close to the experimental single ion-hydration enthalpy for Zn(II). The water molecules of the first sphere form strong hydrogen bonds with water molecules in the second hydration shell because of the strong polarizing effect of the Zn(II) ion. The importance of the second hydration sphere is discussed.  相似文献   

19.
The oxidation of Fe(II) with H2O2 has been measured in NaCl and NaClO4 solutions as a function of pH, temperature T (K) and ionic strength (M, mol-L–1). The rate constants, k (M–1-sec–1), d[Fe(II)]/DT=-k[Fe(II)][2O2]at pH=6.5 have been fitted to equations of the formlog k = log k0+ AI 1/2+BI+CI 1/2/T Where log k0=15.53-3425/T in water; A=–2.3, –1.35; B=0.334, 0.180; and C=391, 235, respectively, for NaCl (=0.09) and NaClO4 ( =0.08). Measurements made in NaCl solutions with added anions yield rates in the order B(OH) 4 >HCO 3 >ClO 4 >Cl>NO 3 >SO 4 2– and are attributed to the relative strength of the interactions of Fe2+ or FeOH+ with these anions. The FeB(OH) 4 + species is more reactive while the FeCO 3 0 , FeCl+, FeNO 3 + and FeSO 4 0 species are less reactive than the FeOH+ ion pair. The general trend is similar to our earlier studies of the oxidation of Fe(II) with O2 except for B(OH) 4 . The effect of pH on the logk was found to be a quadratic function of the concentration of H+ or OH from pH=4 to 8. These results have been attributed to the different rate constants for Fe2+ (k0) and FeOH+ (k1) which are related to the measured k by, k=k0Fe + k1FeOH, where i is the molar fraction of species i. The rates increase due to the greater reactivity of FeOH+ compared to Fe2+. k0 is independent of composition and ionic strength but k1 is a function of ionic strength and composition due to the interactions of FeOH+ with various anions.  相似文献   

20.
The first purely alkoxide-based sol-gel route to nano-phase powders and thin films of perovskite La0.75Sr0.25MnO3 is described. The phase and microstructure evolution on heat treatment of free gel films to form the target nano-phase oxide were investigated by TGA, IR spectroscopy, powder XRD, SEM and TEM-EDS. The xerogel consisted of a hydrated oxo-carbonate, without remaining alkoxo groups or solvent. Heating at 5°C·min–1 decomposed the carbonate groups and yielded the pure perovskite La0.75Sr0.25MnO3 at 760°C. The cell dimensions were virtually unchanged from the first observation of perovskite at 680°C, to 1000°C, 4 h. The monoclinic cell of La0.75Sr0.25MnO3 obtained at 1000°C, 4 h, had the dimensions a = 5.475(1), b = 5.504(2), c = 7.771(1) Å, = 90.50(2), fitting the literature data quite well. Crack-free, homogenous, 150 nm thick La0.75Sr0.25MnO3 films were prepared by spin-coating Si/SiO2/TiO2/Pt and polycrystalline -Al2O3 substrates with a 0.6 M alkoxide solution, followed by heating at 5°C·min–1 to 800°C, 30 min.  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