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 15 毫秒
1.
Equations previously developed and widely applied to the thermodynamic properties of strong electrolytes are extended to solutions involving a dissociation equilibrium. Excellent agreement is obtained with the data for pure phosphoric acid to 6M and for phosphate buffer solutions. The parameters of the strong electrolyte components of the buffer solutions are taken from other work, and the remaining parameters for H+, H2PO 4 , and H3PO4 are evaluated, including a pK of 2.146. The present method avoids ambiguities which formerly arose in treating weak acids with as small pK as this.  相似文献   

2.
The Pitzer ion-interaction model for the thermodynamic properties of a weakly dissociating neutral solute is extended to include solution densities and compressibilities. Excellent agreement is obtained with the literature data for pure aqueous phosphoric acid to 8m. The first and second pressure derivatives of the interaction parameters for H+, H2PO 4 ? , and H3PO4 are evaluated, in addition to the limiting partial molar volumes and adiabatic compressibilities of the ionized and unionized acids. The present method requires no additional data or extraneous assumption regarding the properties of the free electrolyte in extrapolating to infinite dilution. The quantities evaluated are used to estimate the pressure dependence of the ionization constant, activity coefficients, and speciation to 1 kbar. The estimated effect of pressure on aqueous phosphoric acid ionization is in excellent agreement with experimental data.  相似文献   

3.
4.
Proton transport in H3PO4‐ and H2SO4‐blended polybenzimidazoles (PBIs) has been studied with both temperature‐ and pressure‐dependent dielectric spectroscopy. The influences of the acid concentration and temperature on the relative conductance and activation volume are discussed. An Arrhenius relation is used to model the temperature‐dependent conductivity at a constant acid content. The logarithm of the relative conductance for PBI blended with H3PO4 decreases linearly with increasing pressure. As the temperature increases, the activation volume becomes smaller for PBI blended with H3PO4. It is proposed that proton transport in acid‐blended PBI is mainly controlled by proton hopping and diffusion rather than a mechanism mediated by the segmental motions in the polymer. The conductivities of PBIs blended with H3PO4 and H2SO4 are compared. At a 1.45 molar acid doping concentration, the former has the higher conductivity. With water, the conductivity of H3PO4‐blended PBI increases significantly. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 663–669, 2002; DOI 10.1002/polb.10132  相似文献   

5.
Synthesis and Crystal Structure of K2(HSO4)(H2PO4), K4(HSO4)3(H2PO4), and Na(HSO4)(H3PO4) Mixed hydrogen sulfate phosphates K2(HSO4)(H2PO4), K4(HSO4)3(H2PO4) and Na(HSO4)(H3PO4) were synthesized and characterized by X‐ray single crystal analysis. In case of K2(HSO4)(H2PO4) neutron powder diffraction was used additionally. For this compound an unknown supercell was found. According to X‐ray crystal structure analysis, the compounds have the following crystal data: K2(HSO4)(H2PO4) (T = 298 K), monoclinic, space group P 21/c, a = 11.150(4) Å, b = 7.371(2) Å, c = 9.436(3) Å, β = 92.29(3)°, V = 774.9(4) Å3, Z = 4, R1 = 0.039; K4(HSO4)3(H2PO4) (T = 298 K), triclinic, space group P 1, a = 7.217(8) Å, b = 7.521(9) Å, c = 7.574(8) Å, α = 71.52(1)°, β = 88.28(1)°, γ = 86.20(1)°, V = 389.1(8)Å3, Z = 1, R1 = 0.031; Na(HSO4)(H3PO4) (T = 298 K), monoclinic, space group P 21, a = 5.449(1) Å, b = 6.832(1) Å, c = 8.718(2) Å, β = 95.88(3)°, V = 322.8(1) Å3, Z = 2, R1 = 0,032. The metal atoms are coordinated by 8 or 9 oxygen atoms. The structure of K2(HSO4)(H2PO4) is characterized by hydrogen bonded chains of mixed HnS/PO4 tetrahedra. In the structure of K4(HSO4)3(H2PO4), there are dimers of HnS/PO4 tetrahedra, which are further connected to chains. Additional HSO4 tetrahedra are linked to these chains. In the structure of Na(HSO4)(H3PO4) the HSO4 tetrahedra and H3PO4 molecules form layers by hydrogen bonds.  相似文献   

6.
7.
The anodic polarization behaviour of Al-Sn alloy (5.6% Sn) was studied in aerated 1, 1.5 and 2 M H3PO4 acid solutions using potentiodynamic and potentiostatic techniques. Anodic behaviour of Al and Sn metals was investigated for comparison. The results show that the alloy exhibits active-passive transition behaviour. The passivation of the examined alloy is due to the formation of oxide film for both Al and Sn incorporated with their phosphates. In general, at constant concentration of PO4 3− ion, the passive current density (I pass) is decreased with increase of pH in the range 2.5–5. Further, the influence of Cl ions on the passivity of the given alloy was studied. It was found that the aggressive effect of Cl ions on the passive film is inhibited with increase in phosphate concentration and pH. However, the addition of organic compounds (2- and 3-picoline and 2-aminopyridine) shows that only 2-aminopyridine inhibited the attack of Cl ions. Received: 24 October 1997 / Accepted: 5 February 1998  相似文献   

8.
The viscosity, density, and electrical conductivity of H3PO4-dimethylformamide solutions and the enthalpies of solution and mixing of the components were measured in the entire composition range in the interval 25-65°C. It was concluded that phosphoric acid dissociates in dimethylformamide and that complexes of the presumed composition 2H3PO4·DMF are formed.  相似文献   

9.
10.
采用等温溶解法测定了偏钒酸铵(NH4VO3)在NH4H2PO4-H2O和(NH4)3PO4-H2O体系中T = 298.15-328.15 K时的溶解度以及溶液的密度和pH值。结果表明, NH4VO3的溶解度随着(NH4)3PO4或NH4H2PO4溶液浓度的增大,先降低后升高,这是由于同离子效应、化学反应平衡及离子活度的共同作用。比较T = 298.15K时, NH4VO3分别在NH4H2PO4-H2O、(NH4)2HPO4-H2O和(NH4)3PO4-H2O体系中溶解度,发现在相同的磷酸盐浓度下, NH4VO3的溶解度在NH4H2PO4-H2O体系中最大,在(NH4)3PO4-H2O体系中居中,在(NH4)2HPO4-H2O体系中最小。进一步地,在T = 298.15 K和磷酸盐浓度C = 0.5 mol·kg-1时,结合pH值和反应溶度积常数KSP等计算三个体系中的平均离子活度系数(γ±),发现γ±值在(NH4)2HPO4-H2O体系中最大,在(NH4)3PO4-H2O体系中居中,在NH4H2PO4-H2O体系中最小,与溶解度规律一致。  相似文献   

11.
Abstract

A study of several factors has been carried out in order to determine their influence on rare earth phosphates precipitation from H3PO4 solutions obtained after the treatment of the Kola phosphate rock.  相似文献   

12.
13.
Samples of lignocellulosic material, stem of date palm (Phoenix dactylifera), were carbonized at different temperatures (400–600 °C) to investigate the effects of their impregnation with aqueous solution of either phosphoric acid (85 wt%) or potassium hydroxide (3 wt%). The products were characterized using BET nitrogen adsorption, helium pycnometry, Scanning Electron Microscopy (SEM) and oil adsorption from oil–water emulsion (oil viscosity, 60 mPa s at 25 °C). True densities of the products generally increased with increase in carbonization temperature. Impregnated samples (acid/base) showed wider differences in densities at 400 (1.978/1.375 g/cm3) than at 600 °C (1.955/2.010 g/cm3). Without impregnation, the sample carbonized at 600 °C showed higher density of 2.190 g/cm3. This sample has impervious surface with BET surface area of 124 m2/g. Acid-impregnated sample carbonized at 500 °C has the highest surface area of 1100 m2/g and most regular pores as evidenced by SEM micrographs. The amounts of oil adsorbed decreased with increase in carbonization temperature. Without impregnation, sample carbonized at 400 °C exhibited equilibrium adsorption of 4 g/g which decreases to about a half for sample carbonized at 600 °C. Impregnation led to different adsorptive capacities. There are respective increase (48 wt%) and decrease (5 wt%) by the acid- or base-impregnated samples carbonized at 600 °C. This suggests higher occurrence of oil adsorption-enhancing surface functional groups such as carbonyl, carboxyl and phenolic in the former sample.  相似文献   

14.
A 3-D phase diagram of the HPC/H2O/H3PO4 tertiary system against various temperatures was established. Four distinct phases—the completely separated phase (S), the cloudy suspension phase (CS), the liquid crystalline miscible phase (LC), and the isotropically miscible phase (I)—were identified. The S phase shrank as the temperature increased, revealing that the HPC solubility increased with temperature, regardless of the LCST (lower critical solution temperature) characteristic. The addition of H3PO4 suppressed the formation of LC phase. However, as the temperature was raised sharply from 50 to 70?°C, the LC phase could only be maintained at high H3PO4 concentration region; it was a triangular shape, and the top apex of the triangle was the temperature-invariant L* point (HPC/H2O/H3PO4 38/9/53?wt%). The CS phase expanded considerably into the H2O-rich but H3PO4-poor region when the temperature continued to increase over 48?°C. The LCST points of the CS phase that contained 0 and 15?wt% of H3PO4 were 34 and 38?°C, respectively. These CS results demonstrate that H3PO4 suppresses the occurrence of LCST behavior. Additionally, the binodal curve exhibits a weak or even zero dependence of binodal temperature on the HPC concentration at HPC concentrations of less than 30?wt% in a pure water system. A hypothesis concerning the sequential desorption of water molecules was proposed to explain such behavior.  相似文献   

15.
The composite electrolytes of composition (1 ? x)TaH(PO4)2 · 2H2O/xSiO2 (x = 0.2–0.4) are synthesized, and their transport properties are characterized over a wide temperature range. Doping with highly dispersed silica only insignificantly changes the proton conductivity of tantalum hydrogenphosphate below 370 K; above 820 K, the conductivity increases. The evolution of the phase composition of TaH(PO4)2 · 2H2O and its base materials during thermolysis is studied.  相似文献   

16.
Abstract

Besides the production of fertilizers, Kola phosphate rock may be used as a source of lanthanides, strontium and fluorine. It implies the necessity to carry out a study of kinetics and mechanisms of the process in order to choose optimal conditions for the realization of the technological scheme [1]. The fluorapatite concentrate used had the following composition: ?Ln2O3 - 0.89; Y2O3 - 0.04; SrO - 2 80; CaO - 45.40; Fe2O3 -0.42; Al2O3 - 0.86; MgO - 0 10; F - 2.80; SiO2 - 1.80; P2 O 5 - 39.40 weight %; molar ratio CaO : P2O5 = 1.5; the content of the apatite - 98.5%. The reaction of H3PO4 with fluorapatite was studied using a laboratory reactor with a stationary layer. The following parameters were varied: H3PO4 concentration (20, 30 and 50 weight % P2O5), temperature (20, 75, 150, 200 and 250°C) and time of contact (1–180 min.). A multimethod approach was used. X-ray diffraction, electron probe microanalysis and paper chromatography were applied to follow the bulk structural aspects of the apatite powders. It was shown that at the first stage of reaction a thin film of calcium monophosphate Ca(H2PO4)2 H2O is deposited on the apatite particles (avg. diameter 150 nm). The reaction is thought to proceed at the interphase solid/liquid and its kinetics may be described by the equation  相似文献   

17.
A simple dual vessel in situ analyte distillation (IAD) system has been developed for suppressed ion chromatographic determination of chloride and fluoride ions in complex matrices. In IAD system, water vapours generated from the outer vessel reacts with sulfuric acid generating heat, thus favouring the quantitative distillation of chloride and fluoride within 30 min on water bath temperature (approximately 80 degrees C). The distilled analytes, as their respective acids in water, were directly injected into an ion-chromatograph. This newly developed method has been applied for analysis of trace impurities in H2SO4, H3PO4 and H3BO3. The detection limits for chloride is 8, 80 and 70ppb (w/w) for H2SO4, H3PO4 and H3BO3, respectively. For fluoride the detection limits are 6 and 60 ppb (w/w) for H2SO4 and H3PO4, respectively. The recovery of spikes for both the analytes ranged between 87 and 100%.  相似文献   

18.
Conditions of synthesis of poly(ethylene phosphates) in reaction of H3PO4 with HOCH2CH2OH (EG), the actual path of polycondensation, and structure of the obtained polymers (mostly oligomers) and kinetics of reaction are described. Preliminary kinetic information, based on the comparison of the MALDI‐TOF‐ms and 31P{1H} NMR spectra as a function of conversion is given as well. Because of the dealkylation process fragments derived from di‐ and triethylene glycols are also present in the repeating units. Structures of the end groups (? CH2CH2OH or ? OP(O)(OH)2) depend on the starting ratio of [EG]0/[H3PO4]0, although even at the excess of EG the acidic end groups prevail because of the dealkylation process. In MALDI‐TOF‐ms products with Pn equal up to 21 have been observed. The average polymerization degrees (Pn) are lower and have been calculated from the proportion of the end groups. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 830–843, 2008  相似文献   

19.
20.
考察了以亚磷酸作为磷源合成SAPO-41和AlPO-41时加入过氧化氢或磷酸对生成SAPO-41和AlPO-41的影响.结果表明,向反应体系中加入过氧化氢或者使用磷酸部分替代亚磷酸,都能明显缩短SAPO-41和AlPO-41的晶化时间.此外,使用混合磷源还能够降低模板剂的用量,获得小晶粒晶化产物.  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