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 515 毫秒
1.
Thermodynamic and kinetic studies have revealed conventional mean-ion and convenient single-ion activity coefficients for copper(II), magnesium and sulphate ions and standard exchange rates for either step of the liquid Cu(Hg)/Cu(II) electrode and of the solid Cu/Cu(II) electrode, all in 0.005 mol kg?1 CuSO4+0.005 mol kg?1 H2SO4+(x?0.01) mol kg?1MgSO4 (for x=0.2 to 3) at 25°C. With increasing x, the sulphate activity coefficient steadily falls, the copper(II) and magnesium activity coefficients go through a minimum near x=1, and the standard rates fall. The kinetic changes with x reveal some information on double-layer and ligand effects.  相似文献   

2.
Upper critical solution temperatures (UCSTs) of (water + phenol) systems are reported with 0.1 mol · kg−1 halide salts, carboxylic acids, 1.0% PEG 200 in water, and 0.01 mol · kg−1 surfactants and polynuclear aromatic compounds namely benzene, naphthalene, anthracene, chrysene; and benzene derivatives namely toluene and xylene solutions in phenol. Valence electrons and shell numbers, basicity, –CH3 and –CH2–, hydrophilic, hydrophobic and π conjugated electrons of respective additives have been noted to affect their UCSTs and mutual solubilities. The surfactants decrease the USCTs with higher mutual solubilities due to effective hydrophilic as well as hydrophobic interactions with aqueous and organic phases, respectively. A stronger structure breaking action of 3(-OH) of glycerol outweighs those of the 3(-COO) and 1(-OH) of citric acid and urea does produce almost equal UCSTs as compared to glycerol. A decrease in UCSTs is noted with increasing number of conjugated π electrons of benzene, naphthalene, anthracene and chrysene. In general, dTc/dx2 values of salts for 0.20 to 0.16 mole fractions of phenol are found positive while for 0.055 to 0.052 mole fractions, the negative.  相似文献   

3.
The values of the thermodynamic second dissociation constant, pK 2, and related thermodynamic quantities of N-(2-hydroxyethyl)piperazine-N′-2-hydroxypropanesulfonic acid (HEPPSO) have already been reported from 5 to 55?°C, including 37?°C, by the emf method. This paper reports the results for the pH of one chloride-free buffer solution containing the composition: (a) HEPPSO (0.08 mol?kg?1)+NaHEPPSO (0.04 mol?kg?1). The remaining seventeen buffer solutions contain a saline medium of ionic strength I=0.16 mol?kg?1, matching closely that of physiological fluids. Conventional pH values, denoted as pa H, for all eighteen buffer solutions from 5 to 55?°C have been calculated. The operational pH values, designated as pH, with residual liquid-junction corrections for five buffer solutions, one without NaCl, and four with buffer solutions in saline media of I=0.16 mol?kg?1 are recommended as pH standards in the range of physiological application. These are based on the NBS/NIST standard scale for pH measurements.  相似文献   

4.
Platinized platinum based hydrogen gas electrodes, Pt(Pt)|H2(g)|H+(aq), and silver‐silver chloride electrodes, Ag|AgCl|Cl? (aq), make up the Harned cell, without transfer, working in the potentiometric mode at Cl? concentrations and ionic strengths, I, below 0.1 mol kg?1, for assigning primary pH values to reference pH buffer solutions. This work reports on experiments performed at higher I and Cl? solutions up to 0.7 mol kg?1, aiming at addressing seawater conditions with results of equally high quality. In the course of measurements, the occasional occurrence of highly unstable potentials denoted electrode malfunction; Pt metal surfaces observed by SEM/EDS and XRD exhibit strong Ag and Cl peaks corresponding to the presence of AgCl crystals deposited at both surfaces.  相似文献   

5.

Upper critical solution temperatures (UCST) of water‐phenol systems are reported with 0.1 mol kg?1 halide salts, carboxylic acids, 1.0% PEG 200 in water, and 0.01 mol kg?1 surfactants and polynuclear aromatic compounds namely benzene, naphthalene, anthracene, chrysene; and benzene derivatives solutions in phenol. The valence electrons and shell numbers, bascity, ‐CH3 and ‐CH2‐, hydrophilic, hydrophobic and π conjugated electrons of respective additives have been noted to affect the UCST values and mutual solubilities of the water and phenol. The surfactants decrease the UCST values with higher mutual solubilities due to effective hydrophilic as well as hydrophobic interactions with aqueous and organic phases, respectively. The stronger structure breaking action of the 3(‐OH) of the glycerol outweighs than those of the 3(‐COO?) and 1(‐OH) of the citric acid and the urea does produce almost equal UCST values as compared to glycerol. A decrease in the UCST values is noted with number of conjugated π electrons of the benzene, naphthalene, anthracene, and chrysene. In general, the dTc/dx2 values of salts for 0.20–0.16 mole fractions of phenol are found positive while for 0.055–0.052 mole fractions, the negative.  相似文献   

6.
Conventional residual liquid junction potentials were measured between NBS 1:1 phosphate buffer and various dilute solutions with ionic strengths of 1-100 mmol kg?1, using 3.5 mol kg?1 potassium chloride as the bridge electrolyte. All junctions were of the free-diffusion type, formed within a 1-mm capillary. The conventional residual junction potential was indistinguishable from zero (± 0.5 mV), for dilute solutions of primary reference standards, for which calculations based on the Henderson equation predict values of 0.8–1.3 mV. Solutions of Tris-HCl (20–100 mmol kg?1) and dilute HCl (1–50 mmol kg?1) had appreciable conventional residual liquid junction potentials (1.1–3.3 mV). The experimental values for dilute HCl solutions compare favourably with those calculated with the Henderson equation. These results suggest that the pH of hard waters, measured using the NBS pH scale, will be a good approximation (within 0.01 pH) to paH, so that pH can be interpreted in terms of the activity of the hydrogen ions. Further measurements are required to ascertain whether this approximation is true for soft and acidic waters.  相似文献   

7.
The low-lying singlet and triplet states of H2CBe and HCBeH are examined using ab inito molecular orbital theory. In agreement with earlier results, the lowest-lying structure of H2CBe has C2v symmetry and is a triplet with one π electron (3 B1). The results presented here suggest that the lowest-energy singlet structure is the (1B1) open-shell singlet, also with C2v symmetry, at least 2.5 kcal/mol higher in energy. The singlet C2v structure with two π electrons (1A1) is 15.9 kcal/mol higher than 3B1. All of these structures are bound with respect to the ground state of methylene and the beryllium atom. In HCBeH, linear equilibrium geometries are found for the triplet (3Σ) and singlet (1Δ) states. The triplet is more stable than the singlet (1Δ) by 35.4 kcal/mol, and is only 2.9 kcal/mol higher in energy than triplet H2 CBe. Since the transition structure connecting these two triplet molecules is found to be 50.2 kcal/mol higher in energy than H2 CBe, both triplet equilibrium species might exist independently. The harmonic vibrational frequencies of all structures are also reported.  相似文献   

8.
Densities of L-serine, L-isoleucine, L-glutamine in 1.5 mol kg?1 aqueous NaCl, and NaNO3 solutions have been measured for several molal concentrations of amino acids at temperatures from 298.15 to 323.15 K. The partial molar volumes (? v 0 ) of L-serine, L-isoleucine, and L-glutamine in 1.5 mol kg?1 aqueous NaCl/NaNO3 solutions have been computed using density data. The transfer partial molar volumes (Δtr? v 0 ) of L-serine, L-isoleucine, and L-glutamine from water to 1.5 mol kg?1 aqueous NaCl/1.5 mol kg?1 aqueous NaNO3 solutions have been determined at 298.15 K. The trends of variation of ? v 0 and Δtr? v 0 with change in temperature have been discussed in terms of ion-ion, ion-hydrophilic, and ion-hydrophobic interactions operative in solutions.  相似文献   

9.
Isopiestic vapor-pressure measurements were made for Li2SO4(aq) from 0.1069 to 2.8190 mol?kg?1 at 298.15 K, and from 0.1148 to 2.7969 mol?kg?1 at 323.15 K, with NaCl(aq) as the reference standard. Published thermodynamic data for this system were reviewed, recalculated for consistency, and critically assessed. The present results and the more reliable published results were used to evaluate the parameters of an extended version of Pitzer’s ion-interaction model with an ionic-strength dependent third-virial coefficient, as well as those of the standard Pitzer model, for the osmotic and activity coefficients at both temperatures. Published enthalpies of dilution at 298.15 K were also analyzed to yield the parameters of the ion-interaction models for the relative apparent molar enthalpies of dilution. The resulting models at 298.15 K are valid to the saturated solution molality of the thermodynamically stable phase Li2SO4?H2O(cr). Solubilities of Li2SO4?H2O(cr) at 298.15 K were assessed and the selected value of m(sat.)=3.13±0.04 mol?kg?1 was used to evaluate the thermodynamic solubility product K s(Li2SO4?H2O, cr, 298.15 K) = (2.62±0.19) and a CODATA-compatible standard molar Gibbs energy of formation Δf G m o (Li2SO4?H2O, cr, 298.15 K) = ?(1564.6±0.5) kJ?mol?1.  相似文献   

10.
Shake-up satellites corresponding to π* → π transitions accompanying C1s core ionizations have been studied for polystyrene and a series of alkane–styrene copolymers. It is shown that the shake-up intensities are additive in nature and give a direct measure of copolymer compositions.  相似文献   

11.
Ca2+ complexation by both sulfate and selenate ligands was studied by CE. The species were observed to give a unique retention peak as a result of a fast equilibrium between the free ions and the complexes. The change in the corresponding retention time was interpreted with respect to the equilibrium constant of the complexation reaction. The results confirmed the formation of CaSO4(aq) and CaSeO4(aq) under our experimental conditions. The formation data were derived from the series of measurements carried out at about 15, 25, 35, 45 and 55°C in 0.1 mol/L NaNO3 ionic strength solutions, and in 0.5 and 1.0 mol/L NaNO3 ionic strength solutions at 25°C. Using a constant enthalpy of reaction enabled to fit all the experimental data in a 0.1 mol/L medium, leading to the thermodynamic parameters: ΔrG0.1M(25°C)=?(7.59±0.23) kJ/mol, ΔrH0.1 M=5.57±0.80 kJ/mol, and ΔrS0.1 M(25°C)=44.0±3.0 J mol?1 K?1 for CaSO4(aq) and ΔrG0.1 M(25°C)=?(6.66±0.23) kJ/mol, ΔrH0.1 M=6.45±0.73 kJ/mol, and ΔrS0.1 M(25°C)=44.0±3.0 J mol?1 K?1 for CaSeO4(aq). Both formation reactions were found to be endothermic and entropy driven. CaSO4(aq) appears to be more stable than CaSeO4(aq) by 0.93 kJ/mol under these experimental conditions, which correlates with the difference of acidity of the anions as expected for interactions between hard acids and hard bases according to the hard and soft acids and bases theory. The effect of the ionic medium on the formation constants was successfully treated using the Specific ion Interaction Theory, leading to significantly different binary coefficients mol/kg?1 and mol/kg?1  相似文献   

12.
Molten potassium tetrachlorogallate and potassium tetraiodogallate were studied in terms of halogenoacidity, based on X? ion-exchange. Titration of KX solution with GaX3 were achieved and characterized by the shift of cathodic voltammetric curves. Autodissociation constants Ki,X/mol2 kg?1 were determined: ?log Ki,Cl=4.25±0.05 and ?log Ki,I=2.6±0.05, as well as the solubility values of KX: 0.41±0.02 and 0.80±0.02 mol kg?1 for KCl and KI respectively.  相似文献   

13.
The performance of free-diffusion liquid junctions formed in a capillary between saturated potassium chloride solution and a range of solutions with ionic strengths varying from 10?5 to 0.5 mol kg?1 is described. Precision, response time and noise associated with convection were adversely affected at ionic strengths less than 10?3 mol kg?1. Increasing the bore of the capillary from 0.5 mm to 3 mm also had a detrimental effect. Capillaries with 0.5-mm bore performed optimally with errors <0.2 mV for solutions with ionic strength > 10?4 mol kg?1, but deteriorating to 1 mV for 10?5 mol kg?1 solutions. Equivalent errors if free-diffusion junctions were used in pH measurements would be 0.003 and 0.017 pH, indicating that, given well-controlled experimental conditions, it is possible to achieve precise measurements in very dilute solutions.  相似文献   

14.
We calculate the various components of the third-order nonlinear electric susceptibility of the benzene molecule, starting from a CNDO wavefunction. Our theoretical result for the space-fixed nonlinear third-order susceptibility is 6.15 × 10?36 esu, the corresponding experimental value is 1.58 × 10?36 esu. We also estimate the contributions from the σ and π electrons to the total susceptibility.  相似文献   

15.
In this work, hybrid porous Co3O4–CeO2 hollow polyhedrons have been successfully obtained via a simple cation‐exchange route followed by heat treatment. In the synthesis process, ZIF‐67 polyhedron frameworks are firstly prepared, which not only serve as a host for the exchanged Ce3+ ions but also act as the template for the synthesis of hybrid porous Co3O4–CeO2 hollow polyhedrons. When utilized as electrode materials for supercapacitors, the hybrid porous Co3O4–CeO2 hollow polyhedrons delivered a large specific capacitance of 1288.3 F g?1 at 2.5 A g?1 and a remarkable long lifespan cycling stability (<3.3 % loss after 6000 cycles). Furthermore, an asymmetric supercapacitor (ASC) device based on hybrid porous Co3O4–CeO2 hollow polyhedrons was assembled. The ASC device possesses an energy density of 54.9 W h kg?1, which can be retained to 44.2 W h kg?1 even at a power density of 5100 W kg?1, indicating its promising application in electrochemical energy storage. More importantly, we believe that the present route is a simple and versatile strategy for the preparation of other hybrid metal oxides with desired structures, chemical compositions and applications.  相似文献   

16.
Cu2O nanoparticles (nano-Cu2O) modified glassy carbon electrode (GCE) was fabricated and used to investigate the electrochemical behaviour of 4-nitrophenol (4-NP) by cyclic voltammetry (CV), chronoamperometry (CA), chronocoulometry (CC) and differential pulse voltammetry (DPV). Compared with GCE, a remarkable increase in oxidation peak current was observed. It indicates that nano-Cu2O exhibits remarkable enhancement effect on the electrochemical oxidation of 4-NP. Under the optimised experimental conditions, the oxidation peak currents were propotional to 4-NP concentration in the range from 1.0?×?10?6 to 4.0?×?10?4?mol?L?1 with a detection limit of 5.0?×?10?7?mol?L?1 (S/N?=?3). The fabricated electrode presented good repeatability, stability and anti-interference. Finally, the proposed method was applied to determine 4-NP in water samples. The recoveries for these samples were from 94.60% to 105.5%.  相似文献   

17.
The mechanisms of reactions between CC13OO? radical and quercetin, rutin and epigallocatechin gallate (EGCG) have been studied using pulse radiolytic technique. It is suggested that the electron transfer reaction is the main reaction between CC13OO? radical and rutin, EGCG, but there are two main pathways for the reaction of CC13OO? radical with quercetin, one is the electron transfer reaction, the other is addition reaction. The reaction rate constants were determined. It is proved that quercetin and rutin are better CC13OO? radical scavengers than EGCG.  相似文献   

18.
The complexing of palladium(II) with two biological active reagents: glycine (Gly, HA) and 1-aminoethylidene-1,1-diphosphonic acid (AEDP, H4L) at concentrations of chloride ions (0.15 mol/L) corresponding to physiological levels is studied by means of spectrophotometry, pH potentiometry, and 31P NMR spectroscopy. The formation constants for mixed complexes with compositions of [PdH2LA]? (log?? = 43.7) and [PdHLA]2? (log?? = 39.05) are determined. The both ligands are found to be coordinated to palladium(II) in a bidentant-cyclic manner: through amine nitrogen and the oxygen atom of the carboxyl group (in the case of Gly), or through the phosphonic group (in the case of AEDP). A diagram of the distribution of equilibrium concentrations of the complexes depending on pH is calculated for the system K2[PdCl4]: Gly: AEDP = 1: 1: 1. It is demonstrated that there are complexes with compositions of [PdHLA]2?, [PdA2], and [Pd(HL)2]4? in solutions with $C_{Cl^ - } = 0.15 mol/L$ and pH 6?C7.  相似文献   

19.
An electrochemical method for the determination of carbaryl, after prior oxidation to 1,4-naphthoquinone in natural water and soils is reported. The coulometric oxidation of carbaryl at a platinum electrode was studied using 0.024 mol/L Britton-Robinson buffer (pH 7.0). The reduction of the oxidation product 1,4-naphthoquinone at a dropping mercury electrode was used for the indirect determination of carbaryl after separation on C18 Sep-pak cartridges by differential pulse polarography (detection limits: 0.41 mg L?1 of water and 0.47 mg kg?1 of soil) and directly without separation by adsorptive stripping voltammetry (detection limits: 5 μg L?1 of water and 7 μg kg?1 of soil, for 75 s preconcentration time). Relative errors were lower than 3.7% and relative standard deviations smaller than 4.5%.  相似文献   

20.
The photoelectron (PE.) spectra of azuleno[l, 2, 3-cd]phenalene ( 1 ) and azuleno- [5,6,7-cd]phenalene( 2 ) have been recorded. The first five bands of both compounds could be assigned to transitions corresponding to removal of electrons from 4a2, 6b1, 5b1, 3a2 and 4bl orbitals. This assignment is based mainly on a comparison between the observed ionization potentials and orbital energies calculated in a HMO and a PPP model. The UV./VIS. polarized absorption spectrum of 1 in the region 10000–45000 cm?1 has been measured by means of the stretched film technique. The measurements were performed in polyethylene sheets at 77°K. Several bands could be assigned to π* ← π transitions calculated by a PPP-CI method. A comparison between the electronic structures of 1 and 2 is made by means of a simple HMO diagram.  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