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.
In this study, a new set of data for the equilibrium solubility of carbon dioxide in the amine solvent system that consists of triethanolamine (TEA), piperazine (PZ), and water is presented. Equilibrium solubility values were obtained at T = (313.2, 333.2, and 353.2) K and pressures up to 153 kPa using the vapour-recirculation equilibrium cell. The TEA concentrations in the considered ternary (solvent) mixture were (2 and 3) kmol · m?3 and those of PZ’s were (0.5, 1.0, and 1.5) kmol · m?3. The solubility data (CO2 loading in the amine solution) obtained were correlated as a function of CO2 partial pressure, system temperature, and amine composition via the modified Kent–Eisenberg model. Results showed that the model applied is generally satisfactory in representing the CO2 absorption into mixed aqueous solutions of TEA and PZ.  相似文献   

2.
Solubility data of CO2 in aqueous N-methyldiethanolamine (MDEA) solutions of concentration (2.52, 3.36, and 4.28) kmol/m3 were obtained at temperatures (313, 323, and 343) K and partial pressures ranging from about (30 to 5000) kPa. A thermodynamic model based on extended Debye–Hückel theory was applied to predict and correlate of CO2 solubility in various aqueous amine solutions. The effect of piperazine (PZ) concentration on CO2 loading in MDEA solutions was determined at PZ concentration (0.36, 0.86, and 1.36) kmol/m3. Using experimental data in various temperatures the interaction parameters of activity coefficient model for these systems were determined. The results show the model consistency with experimental and literature data and PZ is beneficial to the CO2 loading. The comparison of results of this study with previous data work shows the wide range of CO2 loading considered in this work and the better agreement of model with experimental data. The average absolute relative deviation percent (δAAD) for all data points were 8.11%.  相似文献   

3.
Some heterogeneous reactions of oxide ion exchange (carbonate ion dissociation and magnesium oxide dissolution) in the molten {KCl + LiCl} eutectic at temperatures of (873, 973 and 1073) K were studied using an electrochemical cell with an oxygen membrane electrode Pt(O2)|ZrO2(Y2O3). The dissociation constant of the CO32− was found to increase with increasing temperature: pK (873 K)=(2.39 ± 0.05); pK (973 K)=(1.81 ± 0.09); pK (1073 K)=(1.53 ± 0.08). Removal of CO2 from the gas above the melt allows the complete transformation of CO32− to O2−. pPMgO values decrease more from (6.99 ± 0.08) to (5.41 ± 0.04). The oxobasicity indices, pI(KCl+LiCl), were calculated from the solubility data to be 3.2 at 873 K, 3.4 at 973 K, and 3.6 at 1073 K. This trend suggests an increase in acidity with increasing temperature of {KCl + LiCl}.  相似文献   

4.
In this work, new experimental results for the (vapour + liquid) equilibrium (VLE) of CO2 in piperazine (PZ)-activated concentrated aqueous 2-amino-2-methyl-1-propanol (AMP) are presented for the temperature range of (303 to 328) K and PZ concentration range of (2 to 8) wt.%, keeping the total amine concentration in the solution at 40% and 50 wt.%. The partial pressures of CO2 are in the range of (0.2 to 1500) kPa. The electrolyte non-random two-liquid (ENRTL) theory has been used to develop the VLE model for the quaternary system (CO2 + AMP + PZ + H2O) to describe the equilibrium behaviour of the solution. The CO2 cyclic capacity of these solvents is determined between the rich and lean CO2 loadings. It is found that the CO2 cyclic capacity increases with the addition of PZ in aqueous AMP and also with the increase in AMP concentration in the aqueous solution. However, solid precipitation has been observed for 50 wt.% total amine concentration below T = 318 K for all relative compositions of AMP and PZ in the solvent at higher CO2 loading. The model results of equilibrium composition, pH of the loaded solution and amine volatility of the mixed solvent system, are also presented.  相似文献   

5.
Oxygen permeation through dense ceramic membranes of perovskite-like SrCo0.9−xFe0.1CrxO3−δ (x = 0.01–0.05), Sr1−xyLnxCoO3−δ(Ln = La, Nd, Sm, Gd; x = 0.30–0.35; y = 0–0.10), SrCo1−xTixO3−δ (x = 0.05–0.20) and LaM1−xNixO3−δ (M = Ga, Co, Fe; x = 0–0.6) was studied. The SrCoO3−δ-based solid solutions with cubic perovskite structure were found to exhibit highest permeation fluxes compared to other membranes. However, high thermal expansion coefficients and interaction with gas species such as carbon dioxide may complicate the employment of SrCoO3−δ membranes for oxygen separation membranes. Alternatively, the LaGa1−xNixO3−δ (x = 0.2–0.5) perovskites, having significant permeation fluxes as well as thermal expansion coefficients in the range of (10.8–11.6) × 10−6 K−1, were demonstrated to be suitable as membrane materials at oxygen pressures from 1 × 10−2 to 2 × 104 Pa. Testing oxygen permeation at oxygen partial pressures of 1–60 atm showed that only oxides with a high oxygen deficiency such as SrCo0.85Ti0.15F3−δ possess sufficient oxygen permeation fluxes. The oxygen permeability of perovskites on the basis of LaGaO3 and LaCoO3−δ was found to be negligible at oxygen pressures above 15 atm, caused by low oxygen vacancy concentration and ionic conductivity of such ceramic materials.  相似文献   

6.
The effects of doping the mixed-conducting (La,Sr)FeO3−δ system with Ce and Nb have been examined for the solid-solution series, La0.5−2xCexSr0.5+xFeO3−δ (x = 0–0.20) and La0.5−2ySr0.5+2yFe1−yNbyO3−δ (y = 0.05–0.10). Mössbauer spectroscopy at 4.1 and 297 K showed that Ce4+ and Nb5+ incorporation suppresses delocalization of p-type electronic charge carriers, whilst oxygen nonstoichiometry of the Ce-containing materials increases. Similar behavior was observed for La0.3Sr0.7Fe0.90Nb0.10O3−δ at 923–1223 K by coulometric titration and thermogravimetry. High-temperature transport properties were studied with Faradaic efficiency (FE), oxygen-permeation, thermopower and total-conductivity measurements in the oxygen partial pressure range 10−5–0.5 atm. The hole conductivity is lower for the Ce- and Nb-containing perovskites, primarily as a result of the lower Fe4+ concentration. Both dopants decrease oxide-ion conductivity but the effect of Nb-doping on ionic transport is moderate and ion-transference numbers are higher with respect to the Nb-free parent phase, 2.2 × 10−3 for La0.3Sr0.7Fe0.9Nb0.1O3−δ cf. 1.3 × 10−3 for La0.5Sr0.5FeO3−δ at 1223 K and atmospheric oxygen pressure. The average thermal expansion coefficients calculated from dilatometric data decrease on doping, varying in the range (19.0–21.2) × 10−6 K−1 at 780–1080 K.  相似文献   

7.
Vapour pressures of water over saturated solutions of magnesium, calcium, nickel and zinc acetates were determined as a function of temperature. The vapour pressures served to evaluate the water activities, osmotic coefficients and molar enthalpies of vaporization. Molar enthalpies of solution of magnesium acetate tetrahydrate,ΔsolHm (T =  294.71K ;m =  0.01 mol · kg  1)  =   (15.65  ±  0.97)kJ · mol  1; calcium acetate,ΔsolHm (T =  297.18K ;m =  0.01 mol · kg  1)  =   (28.15  ±  0.28)kJ · mol  1; zinc acetate dihydrate,ΔsolHm (T =  297.36K ;m =  0.01 mol · kg  1)  =   (22.49  ±  0.90)kJ · mol  1and lead acetate trihydrate,ΔsolHm (T =  297.36K ;m =  0.0086 mol · kg  1)  =  (22.46  ±  0.94)kJ · mol  1, were determined calorimetrically.  相似文献   

8.
(Liquid + liquid) equilibria for {1-benzyl-3-methylimidazolium chloride ([BzMIM]Cl) or 1-hexyl-3-methylimidazolium chloride ([HMIM]Cl) + inorganic salts (potassium phosphate K3PO4, potassium carbonate K2CO3, or dipotassium hydrogen phosphate K2HPO4) + H2O} aqueous biphasic systems (ABSs) are presented at T = 298.15 K. An empirical equation was used to correlate the binodal data. The experimental tie lines were appropriately correlated by the Othmer–Tobias and Brancroft empirical equations. The influence of the selected inorganic salts in the phase segregation was investigated by means the calculated effective excluded volume (EEV) and Setschenow-type equation. The salting-out ability of salts was also evaluated in terms of the Gibbs energy of hydration of salt (ΔGhyd) and assessed with EEV values.  相似文献   

9.
Heat capacity of platinic acid, hydrogen hexahydroxyplatinate(IV)H2Pt(OH)6 , was measured from T =  7 K toT =  310 K by means of adiabatic calorimetry. The standard entropy and the standard Gibbs energy of formation of platinic acid in the crystalline state were determined to be 176.5  ±  3.6 J · K  1· mol  1and   988.8  ±  3.8 kJ · mol  1, respectively.  相似文献   

10.
In this study, trimetal Mg/Ce/Mn oxide-modified diatomaceous earth (DE) was synthesized at optimal conditions. Comparison of the SEM images and the results of EDX analyses of the raw and the modified DE confirmed the surface modification of the raw DE with the trimetal oxide. Groundwater fluoride removal capacity of the sorbent was evaluated by batch method at various defluoridation conditions. At a sorbent dosage of 0.6 g/100 mL (contact time: 60 min, mixing speed of 200 rpm and temperature: 297 K), the fluoride removal was >93% for solutions containing initial fluoride concentration of 10–60 mg/L. Sorbent’s optimum fluoride uptake capacity was 12.63 mg/g at the initial fluoride concentration of 100 mg/L. Fluoride removal was >91% for solutions with initial pH range of ∼4–11 (initial fluoride concentration: 9 mg/L, sorbent dosage: 0.6 g/100 mL). Appraisal of the effect of co-existing anions on fluoride removal showed that CO32− would reduce the amount of fluoride removed from solution, while other anions such as PO43−, NO3 and SO42− had no observable effect. K2SO4 solution was found to be most suitable for regeneration of spent Mg/Ce/Mn oxide-modified DE compared to Na2CO3 and NaOH. The mechanism of fluoride removal at pH > 5.45 (pHpzc = 5.45) occurred by exchange of hydroxyl groups on surface of sorbent with fluoride ions from solution. Sorption data fitted better to Langmuir isotherm and pseudo-second-order model. External diffusion was observed to be the sorption rate limiting factor.  相似文献   

11.
Solubilities of l -glutamic acid, 3-nitrobenzoic acid, p -toluic acid, calcium-l -lactate, calcium gluconate, magnesium- dl -aspartate, and magnesium- l -lactate in water were determined in the temperature range 278 K to 343 K. The apparent molar enthalpies of solution at T =  298.15 K as derived from these solubilities areΔsolHm (l -glutamic acid,msat =  0.0565 mol · kg  1)  =  30.2 kJ · mol  1,ΔsolHm (3-nitrobenzoic acid, m =  0.0188 mol · kg  1)  =  28.1 kJ · mol  1, ΔsolHm( p - toluic acid, m =  0.00267 mol · kg  1)  =  23.9 kJ · mol  1,ΔsolHm (calcium- l -lactate tetrahydrate,m =  0.2902 mol · kg  1)  =  25.8 kJ · mol  1,ΔsolHm (calcium gluconate, m =  0.0806 mol · kg  1)  =  22.1 kJ · mol  1, ΔsolHm(magnesium-dl -aspartate tetrahydrate, m =  0.1469 mol · kg  1)  =  11.5 kJ · mol  1, andΔsolHm (magnesium- l -lactate trihydrate,m =  0.3462 mol · kg  1)  =  3.81 kJ · mol  1.  相似文献   

12.
This report presents a new set of heat capacity data for the system piperazine {(PZ) + 2-amino-2-methyl-1-propanol (AMP) + water (H2O)}, measured using the differential scanning calorimetry technique, over the temperature range 303.2 K to 353.2 K and at fourteen (14) different concentrations in which the water mole fractions, x3’s, were fixed at 0.60, 0.70, 0.80, and 0.90. Heat capacity for the binary system {PZ (1) + AMP (2)} at x1 = 0.05, 0.10, 0.15, and 0.20 were, likewise, measured to generate parameters necessary in the Redlich–Kister-type model, which was used to estimate excess molar heat capacities. Such estimates were then used to predict the values of the molar heat capacity at the corresponding sets of temperature and concentration. The predicted values were subsequently compared against the measured values and the results are satisfactory.  相似文献   

13.
The thermodynamic properties ofZn5(OH)6(CO3)2 , hydrozincite, have been determined by performing solubility and d.s.c. measurements. The solubility constant in aqueous NaClO4media has been measured at temperatures ranging from 288.15 K to 338.15 K at constant ionic strength (I =  1.00 mol · kg  1). Additionally, the dependence of the solubility constant on the ionic strength has been investigated up to I =  3.00 mol · kg  1NaClO4at T =  298.15 K. The standard molar heat capacity Cp, mofunction fromT =  318.15 K to T =  418.15 K, as well as the heat of decomposition of hydrozincite, have been obtained from d.s.c. measurements. All experimental results have been simultaneously evaluated by means of the optimization routine of ChemSage yielding an internally consistent set of thermodynamic data (T =  298.15 K): solubility constant log * Kps 00 =  (9.0  ±  0.1), standard molar Gibbs energy of formationΔfGmo {Zn5(OH)6(CO3)2 }  =  (  3164.6  ±  3.0)kJ · mol  1, standard molar enthalpy of formation ΔfHmo{Zn5(OH)6(CO3)2 }  =  (  3584  ±  15)kJ · mol  1, standard molar entropy Smo{Zn5(OH)6(CO3)2 }  =  (436  ±  50)J · mol  1· K  1and Cp,mo / (J · mol  1· K  1)  =  (119  ±  11)  +  (0.834  ±  0.033)T / K. A three-dimensional predominance diagram is introduced which allows a comprehensive thermodynamic interpretation of phase relations in(Zn2 +  +  H2O  +  CO2) . The axes of this phase diagram correspond to the potential quantities: temperature, partial pressure of carbon dioxide and pH of the aqueous solution. Moreover, it is shown how the stoichiometric composition{n(CO3) / n(Zn)} of the solid compoundsZnCO3 and Zn5(OH)6(CO3)2can be checked by thermodynamically analysing the measured solubility data.  相似文献   

14.
Extreme ultraviolet (EUV) spectroscopy was recorded on microwave discharges of helium with 2% hydrogen. Novel emission lines were observed with energies of q·13.6 eV where q=1,2,3,4,6,7,8,9, or 11 or these lines inelastically scattered by helium atoms wherein 21.2 eV was absorbed in the excitation of He (1s2) to He (1s12p1). These lines were identified as hydrogen transitions to electronic energy levels below the ‘ground’ state corresponding to fractional quantum numbers. Significant line broadening corresponding to an average hydrogen atom temperature of 33–38 eV was observed for helium–hydrogen discharge plasmas; whereas pure hydrogen showed no excessive broadening corresponding to an average hydrogen atom temperature of ≈3 eV. Since a significant increase in H temperature was observed with helium–hydrogen discharge plasmas, and energetic hydrino lines were observed at short wavelengths in the corresponding microwave plasmas that required a very significant reaction rate due to low photon detection efficiency in this region, the power balance was measured on the helium–hydrogen microwave plasmas. With a microwave input power of 30 W, the thermal output power was measured to be at least 300 W corresponding to a reactor temperature rise from room temperature to 900 °C within 90 s, a power density of 30 MW/m3, and an energy balance of about −4×105 kJ/mol H2 compared to the enthalpy of combustion of hydrogen of −241.8 kJ/mol H2.  相似文献   

15.
In the present work, the solubility of CO2 in aqueous solutions of potassium prolinate (KPr) and potassium α-aminobutyrate (KAABA) was measured at temperatures (313.2, 333.2, and 353.2) K and CO2 partial pressures up to 1000 kPa for amino acid salt concentrations: KPr, w = (7.5, 14.5, and 27.4 wt%) and KAABA, w = (6.9, 13.4, and 25.6 wt%). It was found that the CO2 absorption capacities of the studied amino acid salt systems were considerably high and comparable with that of industrially important alkanolamines including monoethanolamine. The CO2 loadings in aqueous potassium α-aminobutyrate at high pressures were also found to be generally higher than the loadings in aqueous potassium prolinate. A modified Kent–Eisenberg model was applied to correlate the CO2 solubility in the amino acid salt solution as function of CO2 partial pressure, temperature, and concentration. The model gave good representation of the (vapour + liquid) equilibrium data obtained for the amino acid salt systems studied, and provided accurate predictions of the solubility.  相似文献   

16.
《Polyhedron》2005,24(16-17):2269-2273
Two ion-pair compounds, consisting of 1-(4′-R-benzyl)pyridinium ([RBzPy]+, R = NO2 (1) and Br (2)) and [Ni(dmit)2] (dmit2− = 2-thioxo-1,3-dithion-4,5-dithiolato), have been synthesized and structurally characterized. The anions of [Ni(dmit)2] stack into dimers, which further construct into two-leg ladder through terminal S⋯S interactions in 1, lateral S⋯S interactions in 2. The weak H-bonding interactions of C–H⋯S were observed in 2, while only weak van de Waals interactions between anion and cations in 1. The magnetic susceptibilities measured in 2–300 K indicate AFM exchange interaction domination both two compounds. A peculiar magnetic transition at ∼100 K was observed in 1. An AFM ordering below ∼11 K was found in 2, and the best fit to magnetic susceptibility above 45 K in this compound, using a dimer model with s = 1/2, give rise to Δ/kB = 36.1 K, zJ = −0.91 K, C = 3.2 × 10−3 emu K mol−1 and χ0 = −4.0 × 10−6 emu mol−1 with g of 2.0 fixed.  相似文献   

17.
In this work, a new set of values for the solubility of carbon dioxide in aqueous mixture containing different concentrations of 2-amino-2-methyl-1-propanol (AMP), a sterically-hindered amine, and piperazine (PZ), an activator, are presented. The results were carefully determined using a 1.0 dm3 stainless steel vapour-recirculation equilibrium cell at T = (313.2, 333.2, and 353.2) K, and pressures up to 152 kPa. The AMP concentrations in the ternary (solvent) mixture were (2 and 3) kmol · m?3; those of PZ’s were (0.5, 1.0, and 1.5) kmol · m?3. The measured equilibrium loading (solubility)/partial pressure pairs at different temperatures and concentration levels were generally consistent with the corresponding values correlated from the Kent–Eisenberg model that has been adapted for the system in the study, where the parameters of the models were determined using the results from this study and relevant data from literature.  相似文献   

18.
《Polyhedron》2002,21(14-15):1481-1484
The formation of complexes between phosphocreatine, H2O3PNHC(NH)N(CH3)CH2CO2H, and the ions Na+, K+ and Mg2+ have been investigated under physiological conditions (aqueous solution, T=37 °C and I=0.25 mol dm−3) by means of 31P NMR spectroscopy. Only 1:1 complexes have been identified. Stability constants have been determined with the aid of the new computer program hypnmr-2000. log10 K values were found to be−0.5(2),−0.3(2) and 1.43(3), respectively. The formation constant for the potassium complex is two orders of magnitude less that the literature value.  相似文献   

19.
Isopiestic vapor-pressure measurements were made for Rb 2SO 4(aq) from molalitym =  (0.16886 to 1.5679 )mol · kg  1atT =  298.15 K and from m =  (0.32902 to 1.2282 )mol · kg  1at T =  323.15 K, and for Cs 2SO4 (aq) from m =  (0.11213 to 3.10815 )mol · kg  1at T =  298.15 K and fromm =  (0.11872 to 3.5095 )mol · kg  1atT =  323.15 K, with NaCl(aq) as the reference standard. Published thermodynamic information for these systems were reviewed and the isopiestic equilibrium molalities and dilution enthalpies were critically assessed and recalculated in a consistent manner. Values of the four parameters of an extended version of Pitzer`s model for osmotic and activity coefficients with an ionic-strength dependent third virial coefficient were evaluated for both systems at both temperatures, as were those of the usual three-parameter Pitzer model. Similarly, parameters of Pitzer`s model for the relative apparent molar enthalpies of dilution were evaluated at T =  298.15 K for both Rb 2SO 4(aq) and Cs 2SO 4(aq) for the more restricted range of m⩽ 0.101 mol · kg  1. Values of the thermodynamic solubility product Ks(Rb2 SO 4, cr, 298.15 K )  =  (0.1392  ±  0.0154) and the CODATA compatible standard molar Gibbs free energy of formationΔfGmo (Rb 2SO 4, cr, 298.15 K )  =   (1316.91  ±  0.59)kJ · mol  1, standard molar enthalpy of formationΔfHmo (Rb 2SO 4, cr, 298.15 K )  =   (1435.07  ±  0.60)kJ · mol  1, and standard molar entropy S mo(Rb2 SO 4, cr, 298.15 K )  =  (199.60  ±  2.88)J · K  1· mol  1were derived. A sample of one of the lots of Rb 2SO 4(s) used for part of our isopiestic measurements was analyzed by ion chromatography, and was found to be contaminated with potassium and cesium in amounts that significantly exceeded the claims of the supplier. In contrast, analysis by ion chromatography of a lot of Cs 2SO 4(s) used for some of our experiments showed it was highly pure.  相似文献   

20.
(Vapour + liquid) equilibrium data (water activity, vapour pressure, osmotic coefficient, and activity coefficient) of binary aqueous solutions of 1-hexyl-3-methylimidazolium chloride ([C6mim][Cl]), methyl potassium malonate, and ethyl potassium malonate and ternary {[C6mim][Cl] + methyl potassium malonate} and {[C6mim][Cl] + ethyl potassium malonate} aqueous solutions were obtained through the isopiestic method at T = 298.15 K. These results reveal that the ionic liquid behaves as surfactant-like and aggregates in aqueous solutions at molality about 0.4 mol · kg−1. The constant water activity lines of all the ternary systems investigated show small negative deviations from the linear isopiestic relation (Zdanovskii–Stokes–Robinson rule) derived using the semi-ideal hydration model. The density and speed of sound measurements were carried out on solutions of methyl potassium malonate and ethyl potassium malonate in water and of [C6mim][Cl] in aqueous solutions of 0.25 mol · kg−1 methyl potassium malonate and ethyl potassium malonate at T = (288.15 to 308.15) K at atmospheric pressure. From the experimental density and speed of sound data, the values of the apparent molar volume, apparent molar isentropic compressibility and excess molar volume were evaluated and from which the infinite dilution apparent molar volume and infinite dilution apparent molar isentropic compressibility were calculated at each temperature. Although, there are no clear differences between the values of the apparent molar volume of [C6mim][Cl] in pure water and in methyl potassium malonate or ethyl potassium malonate aqueous solutions, however, the results show a positive transfer isentropic compressibility of [C6mim][Cl] from pure water to the methyl potassium malonate or ethyl potassium malonate aqueous solutions. The results have been interpreted in terms of the solute–water and solute–solute interactions.  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