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.
The effect of the surface conductance on the -potential of dispersed particles determined from their electrophoretic mobility in nonaqueous electrolyte solutions was considered. The conductivity of dilute quartz suspensions and the electrophoretic mobility of quartz particles in NaBr solutions in butanol-1 and dimethyl sulfoxide as well as in LiBr solutions in acetone at the salt concentration C= 10–5–10–2M were determined by conductometry and microelectrophoresis. The dependences of surface conductivity and -potential on the electrolyte content were calculated by formulas of the Wagner and Henry theories. It was shown that, in the region of dilute solutions, (log C) curves thus obtained significantly differ from corresponding functions calculated by the Smoluchowski equation. At the same time, these dependences agreed closely with the (log C) dependences determined for the same systems by the streaming potential method with allowances for experimental values of the surface conductivity. Using aluminum oxide suspensions in NaBr and HBr ethanol solutions as examples, it was shown that, to obtain correct values of the -potential from electrophoretic mobility of porous particles impregnated with a solution, it is necessary to allow for the bulk conductance of the particles.  相似文献   

2.
The conductivity of dilute quartz suspensions and electrophoretic mobility of quartz particles in solutions with the concentration C = 10–5–10–2 M XBr (X = H, Cs, Na, and Li) and NaOH, as well as in mixed solutions of 10–4 M XBr (X = Cs, Na, and Li) + 10–4–10–2 M HBr and 10–4 M XBr + 10–4–10–2 M XOH (X = Cs, Na, and Li) in ethanol containing 6 vol % of water were measured using conductometry and microelectrophoresis. The values of surface conductivity of quartz were calculated by the Wagner formula and used to calculate zeta potential by the Henry–Booth formula. The resultant dependences (logC) suggest that the value and sign of zeta potential are determined not only by the adsorption of potential-determining ions + and , but also by the competitive specific adsorption of all ions of the aforementioned electrolytes, the adsorption values increasing in a cation series Li+ < Na+ < Cs+ < H+ and an anion series Br < OH. In particular, it is found that the titration of the above suspensions with XOH bases results in the reversal of zeta potential sign from negative to positive at a concentration depending on the adsorption capacity of alkali cation.  相似文献   

3.
The kinetics of the gold dissolution in cyanide solutions is studied at constant values of the coverage () of the gold surface by mercury atoms. The constancy of is ensured by maintaining an identical value of the duration (t) of contact of electrode with solution (after renewing its surface by cutting off a thin surface layer of metal) at a potential of –1.3 V, at which the discharge of mercury ions is limited by their diffusion to the electrode. At t = const kinetic dependences of the gold dissolution process correspond to the Tafel equation. Effective values of exchange current i 0, transfer coefficient , and reaction order by cyanide ions P are determined. With increasing value of their magnitude increases from values 10–5 A cm–2, 0.1, and 0.17 that are characteristic of purely cyanide solutions (composition 0.1 M KCN, 0.1 M KOH, and 0.01 M KAu(CN)2) to i 0 2 × 10–4 A cm–2, = 0.46, and P 1 at t = 270 s. These results are compared with the data obtained earlier during similar investigations in solutions containing thallium, lead, and bismuth. Common and individual features in the behavior of mercury-containing electrolytes are revealed. It is shown that the possible mechanism of the acceleration of the gold dissolution process in the presence of catalytically active atoms, which had been proposed in these works, may be used also for explaining the action of mercury atoms on this process.  相似文献   

4.
Electrosurface properties (the -potential and surface conductivity) of quartz particles in water–ethanol solutions of CsBr, NaBr, and LiBr with concentrations C = 10–5–10–2 M are studied. The (log C) dependences plotted from the results of electrophoretic measurements with allowance made for the particle surface conductivity demonstrate that, when water content in the aforementioned solutions increases from 4 to 40 vol %, the -potential of quartz becomes more negative and the isoelectric point shifts toward higher electrolyte concentrations, which increase in the following series: CsBr < NaBr < LiBr. This shift of the isoelectric point is explained by a decrease in the specific interaction of the alkali metal cations with the quartz surface because of a rise in the degree of their hydration (supersolvation).  相似文献   

5.
The molecular structure of BeBr2 has been investigated by gas-phase electron diffraction at the temperature 800(10) K. The conventional analysis yielded the following values: r g(Be–Br) = 1.944(6)Å, l(Be–Br) = 0.068(4)Å, r g(Br–Br) = 3.848(8)Å, l(Br–Br) = 0.109(3)Å, k(Be–Br) = 1.1(1.1) × 10–5 Å3, (Br–Br) = 2.1(1.0) × 10–5 Å3. Three models of nuclear dynamics were used to simulate the conventional analysis values—infinitesimal vibrations and two models, which take into account the kinematic and dynamic anharmonicity of the bending vibration. All models give similar values of bond angle, amplitudes, and shrinkage, excluding the harmonic model, which yields too low value l(Br–Br). The equilibrium bond distance r e(Be–Br) = 1.932(11) Å was estimated, taking into account the anharmonicity corrections for stretching and bending vibrations and centrifugal distortion.  相似文献   

6.
The complexes (OC)4(CNBu t )ReOs(CO)3(CNBu t )Os(CO)3(CNBu t )Re(CNBu t )(CO)4 (A) and (OC)3(CNBu t )2ReOs(CO)4Os(CO)3(CNBu t )Re(CNBu t )(CO)4 (B) have been isolated in low yield from the reaction of Os(CO)3(CNBu t )2 with Re2(-H)(--C2H3)(CO)8 in hexane at room temperature. Both compounds have approximately linear ReOs2Re chains. The Re–Os lengths are in the range 2.9311(7)–2.952(1) Å the Os–Os lengths are 2.875(1) (A) and 2.8759(7) Å (B).  相似文献   

7.
Summary The species, UO2H3L, UO2H2L2–, UO2HL3–, UO2L4–, UO2(OH)L5– and UO2(OH)2L6– are found in the equilibria between uranyl ions and 3,3-bis[N,N-di(carboxymethyl)-aminomethyl]-o-cresolsulphonphthalein (H6L; xylenol orange; dcac) in aqueous solution. The equilibria have been studied by the potentiometric method at 25° and at an ionic strength of 0.1M (KNO3). New algebraic equations have been employed to evaluate the equilibrium constants.  相似文献   

8.
Quartz microgravimetry is used to determine the ratio between coefficients of mass transfer (1 : 0.54 : 0.48), which characterizes relative values of rates of diffusion of hydroxy complexes of thallium, lead, and bismuth in alkaline solutions. The ratio is used when refining the condition under which on a renewable electrode in solutions containing these ions at the concentration c i the electrode coverage by relevant adatoms i with increasing duration of contact of the electrode with solution reaches constant values: (c Tl t) = (0.54c Pb t) = (0.48c Bi t). Measured are i,t curves on a renewable gold electrode at E = const in solutions containing 0.1 M KCN, 0.1 M KOH, 0.01 M KAu(CN)2, and 8 × 10–6 M compounds of thallium or 1.5 × 10–5 M, lead, or 1.6 × 10–5 M, bismuth. Shown is that at (, E) = const the currents of dissolution of gold in these solutions increase in the series Tl < Pb < Bi, which evidences an increase in this series of the catalytic activity of adatoms of these metals. Shown is that at = const the catalytic action of adatoms of thallium and bismuth has an approximately additive character. The obtained data are analyzed with allowance made for the explanation offered earlier for the catalytic effect of adatoms on the anodic dissolution of gold based on the hypothesis about the shift of the potential of the free zero charge in the negative direction after substituting a metal atom for chemisorbed cyanide ions.  相似文献   

9.
A new CrIII complex, viz. trans-[Cr(naphprn)(H2O)2]ClO4·2H2O, (1) {naphprn=N,N-trimethylenebis(naphthylideneimine)} has been synthesized and characterized. Aquo ligand substitution of ( 1 ) by the azide ligand gave rise to trans-[Cr(naphprn)(N3)(H2O)], ( 2 ). Irradiation of ( 2) in DMF gave nitrido(naphprn)chromium(V) ( 3 ). Solutions of ( 3 ) showed e.p.r. spectra at room temperature (g iso=1.9865). The i.r. spectra showed disappearance of a band at 2067cm–1 showing the breakdown of NN and appearance of a new band at 1074cm–1which is assigned to the Cr14N, stretching frequency indicating the formation of a nitrido chromium(V) complex, ( 3). The u.v.–vis. spectrum of complex ( 3 ) exhibited a d–d band maximum at 553nm (=120M–1cm–1). The rate of formation of ( 2 ) was found to be 5.0×10–3M–1s–1 in an aqueous acidic medium at [CrIII]=0.5×10–3M; [N3 ]=0.01–0.15M; [H+]=0.001M and I=0.2M (LiClO4). The rate of photo-decomposition of ( 2 ) to give rise to ( 3 ) was found to be 0.15×10–3s–1 in DMF.  相似文献   

10.
The photo-absorbing, basic sensor, 4-nitroaniline, has been used to determine theequilibrium constant for solvent reorganization around the proton in mixtures ofvarying composition of water with acetic acid. In all the mixtures used, theself-ionization of the acetic acid was suppressed. In contrast to mixtures of waterwith the related ethanol or acetone, this equilibrium is shifted more toward thewater-solvated species as the mole fraction x 2 of the cosolvent increases. TheGibbs energy of transfer of protons from water into the mixture G o t (H+) can bederived with the aid of this equilibrium constant for the solvent reorganization.Using G o t (H+), G o t (i) for i denoting anions and other cations can be evaluated.In comparison the G o t (i) for cations have lower negative values than when eitherethanol or acetone is added to water. Correspondingly, for halide anions, thepositive G o t (i) with added acetic acid are rather less than is found with eitherethanol or acetone added. The influence on the ion-solvent interaction of bothelectron withdrawing hydroxy and carbonyl groups in acetic acid may beresponsible for this. Although G o t (i) for C10 4 and Re0 4 are also positive, both picrateions and OH give negative values with acetic acid added to water. With picrateions, the hydrophobic effect of the carbon ring produces stabilization in themixture relative to water. With OH, complete conversion to acetate anionsoccurs. As is found with other cosolvents, the contribution of the charge onacetate anion to G o t (CH3COO) is found to increase as x 2 rises. The aciddissociation constant K a for acetic acid is found to decrease slowly as x 2 rises to0.5, followed by a rapid decrease for x 2 greater than 0.7 where dimerization ofacetic acid occurs.  相似文献   

11.
Solid solutions of the structure -K4P2O7, which have the composition (K1 – x Rb x )3.8M0.1P2O7(M = Ca, Sr, Cd, Ba) and (K1 – x Rb x )4 – 2y Ca y P2O7(y= 0.025–0.20), are synthesized. The solutions' electroconductivity is studied in the temperature range 400–600°C. The concentration dependences of the conductivity and the activation energy for conduction have extremums at x 0.6, which points to the polyalkaline effect. The effect increases with decreasing radius of cation M2+and with increasing concentration of calcium cations, which is explained by an increased energy nonequivalence of positions in the cation sublattice accessible to alkali ions.  相似文献   

12.
The gold dissolution rate iin solutions containing 0.1 M KOH, 0.1 M KCN, and 2.5 × 10–7to 1.5 × 10–5M TlNO3is studied as a function of potential Eof the electrode whose surface is renewed prior to each experiment, the TlNO3concentration c, and the time tof the electrode contact with solution. At cexceeding 0.5 × 10–5M and t 0, the rate is 1.5–2 times that at c= 0. Initial portions of ivs. tcurves in the absence and presence of TlNO3coincide only at cbelow 10–6M. Potentiostatic and potentiodynamic measurements show that, at positive E, only small coverages of the electrode surface with thallium are obtained, which make no impact on iat E< 0 and heavily increase it at 0 < E< 0.3 V. The discovered effects are attributed to the formation, during the adsorption of oxidized thallium forms, of dipoles comprising thallium adions and gold atoms. Presumably, the dipoles face the gold with their negative ends and make the potential of zero free charge more negative.  相似文献   

13.
The cluster Os7(CO)20(CNBu t ) (1) has been prepared in 25% yield by the reaction of Os6(CO)18 with Me3NO and Os(CO)4(CNBu t ) at –78°C. The crystal structure of 1 reveals the expected capped octahedral arrangement of metal atoms with the noncarbonyl ligand attached to the capping Os atom. The OsOs lengths in the two independent molecules in the unit cell are in the range 2.823(1)–2.922(1) Å, with the longer bonds associated with the Os3 triangle farthest from the capping Os atom. The 13C NMR spectrum of 1 in solution at room temperature has a 3:3:1 pattern that is consistent with rotation of the individual Os(CO)2(L) (L=CO or CNBu t ) groups in the cluster. This in turn supports the idea that the capping Os(CO)2(CNBu t ) unit binds to the central Os6 via a centrally directed MO plus two tangential molecular orbitals.  相似文献   

14.
Solutions of chitosan with molecular weight (MW) of 20000, 9600, and 3700 Da are studied conductimetrically and viscosimetrically. The dependence of solution conductivity on the chitosan concentration begins to deviate from linearity simultaneously with an abrupt increase in the solution viscosity starting from concentrations of 20–30 g l–1. The fraction of free counterions (Cl, CH3COO) in the 0.1 g-equiv l–1 chitosan solutions significantly depends on the sample's MW. The charge is transferred in solutions predominantly by chloride and acetate ions, with the high-MW cation barely contributing to conduction.  相似文献   

15.
Two conformers (chair, boat) of [l-(–)-menthyl)]-[2,2-methylene-bis-(4-methyl-6-tert-butylphenyl)] phosphite ozonide have been obtained by the low temperature ozonization (–80 °C) of [l-(–)-menthyl)]-[2,2-methylene-bis-(4-methyl-6-tert-butylphenyl)] phosphite. It was determined that decomposition of the ozonide is first order with the rate constant logk 0 = (10.92±1.10)–(14.02±1.25)/gq ( = 2.303RT, kcal mol–1), leading to [l-(–)-menthyl)]-[2,2-methylene-bis-(4-methyl-6-tert-butylphenyl)] phosphate and oxygen (including singlet oxygen). Conformational transitions (chair-boat) for [l-(–)-menthyl)]-[2,2-methylene-bis-(4-methyl-6-tert-butylphenyl)] phosphate have been registered by31P NMR spectroscopy.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1758–1761, October, 1994.This work was supported by Russian Foundation for Basic Research (Project No. 93-03-532l).  相似文献   

16.
Reactive extraction separation of binary amino acids from water using a microporous hollow fiber has been studied, in which the acidic extractant di(2-ethylhexyl)phosphoric acid (D2EHPA) was selected as an active carrier dissolved in kerosene. l-Phenylalanine (Phe) was extracted from an aqueous solution through the shell side of module to the organic phase through the lumen of fiber in the extraction module, in which l-Phe was then back-extracted to stripping phase in stripping module. Experiments were conducted as a function of the initial feed concentration of equimolar Phe and l-aspartic acid (l-Asp) (5 mol/m3), feed pH (3–5), the carrier concentration (0.1–0.5 mol/dm3), and stripping acidity (0.1–2 mol/dm3). The effect of process variables on the separation factor of Phe/Asp and the possible transport resistances including aqueous-layer diffusion, membrane diffusion, organic-layer, and interfacial chemical reaction were quantitatively studied and discussed. The high separation factor (β) of Phe/Asp was obtained to be 18.5 at feed pH 5 and 2 mol/dm3 of strip solution (HCl). The extraction and stripping processes appear to rely on pH dependence of the distribution coefficient of amino acids in reactive extraction system. The separation factor (β) was enhanced in hollow fiber membrane (HFM) process compared with conventional solvent process, which was a result of the counter transport of hydrogen ions.  相似文献   

17.
Foam films and wetting films on quartz formed from aqueous solutions of cetyltrimethylammonium bromide (CTAB) are investigated in a wide range of surfactant concentrations in the presence of background electrolyte (5 × 10–4 mol dm–3 NaCl). Foam and wetting films are convenient models for the study of symmetric (free thin liquid films) and asymmetric (thin liquid films on solid substrate) films with the same air/solution interface. Microinterferometric methods of assessment of foam and wetting films are used which allow precise determination of the film thickness. Determined are the values of the potential 0 of the diffuse electrical layer at the solution/air interface (applying the method of equilibrium foam films) and the potential 1 at the solution/quartz interface (applying the method of capillary electrokinetics). These values are used to analyze the stability of the films studied in terms of the DLVO theory. A conclusion drawn is that both kinds of films studied are stabilized by electrostatic interaction forces. It is shown that with increasing CTAB concentration, a charge reversal occurs at both the solution/air and solution/quartz interfaces which determines the stability/instability conditions of the foam and wetting films. Concentration ranges where both kinds of films produce stable (equilibrium) films are found. There are also concentration ranges where the films either rupture or are metastable (quasi-equilibrium). The CTAB concentration ranges, which provide the formation of unstable (rupturing and metastable) and stable films, are different for symmetric (foam) and asymmetric (wetting) thin liquid films. It is only at high CTAB concentrations (higher that >2 × 10–4 mol dm–3) that both cases render formation of stable equilibrium films. These studies give direct experimental indications that the electrostatic interactions between identical or different interfaces can differ when the surfactant concentration is varied.  相似文献   

18.
Measurements are reported on the formation of negative ions in O2, O2/Ar and O2/Ne clusters aimed at establishing the mechanisms of anion formation and the role of inelastic electron scattering by the cluster constituents on negative ion formation in clusters. In the case of pure O2 clusters the main anions we detected are of two types: O(O2) n0 and (O2) n 1– . The yields of O(O2) n showed maxima at 6.3, 8.0 and 14.0 eV and the data suggest O as their precursor; the maxima at 8 and 14 eV are due to the production of O via symmetry forbidden dissociative attachment processes in O2 at these energies which become allowed in clusters. The yields of (O2) n showed a strong maximum at near-zero energy (0.5 eV) and also at 6.3, 8 and 14 eV. With the exception of the near-zero energy resonance, the (O2) n anions at 6.3, 8 and 14 eV are attributed to nondissociative attachment of near-zero energy secondary electrons to O2 clusters. The slow secondary electrons result predominantly from scattering via the O 2 negative ion states of incident electrons with energies in their respective regions. Similar results were obtained for the mixed O2/rare gas clusters except that now a feeble and distinctly structured contribution in the yields of O(O2) n , (O2) n (and Ar(O2) n ) was observed at energies >10 eV. These anions are believed to have the lowest negative ion states of Ar* (Ne*) as their precursors.  相似文献   

19.
The multi-centre integrals of the orbital system n Y lm () exp (–r 2) are evaluated using the Talmi transformation of nuclear shell theory. The integrals are simpler than those of the systems r 2n Y lm(r) exp (–r 2), x l y m z n exp (–r 2), (/x) l (/y) m (/z) n exp (–r 2) and the spherical oscillator functions. The integral types investigated are: overlap, electric dipole transition (momentum operator), kinetic energy, three-centre nuclear attraction, four-centre electronic repulsion, three-centre spin-orbit coupling, and magnetic dipole transition (three-centre integrals of the angular momentum operator).  相似文献   

20.
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   

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