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.
Densities and viscosities of binary mixtures (H2O or D2O) (1) + (DMSO or DMSO-D6)(2) have been measured over the entire mole fraction range; and the excess volumes, excess viscosities, and excess partial molar volumes Vf of the components have been obtained. All systems show negative excess volume Ve at all compositions, values for mixtures containing D2O being more negative than those with H2O byca. 0.03 cm3-mol-1 at x1, = 0.6, where a minimum is observed. The difference between DMSO and DMSO-D6 containing mixtures is negligible. The excess viscosity ηe is always positive and shows a maximum at x1 = 0.65; at this composition, the substitution of H2O with D2O causes an excess viscosity increment ofca. 0.35 mPa-s, while deuteration of DMSO brings about a smaller increase,ca. 0.1 mPa-s. The trend of V 2 E with concentration shows the characteristic features of moderately hydrophobic solutes in water (negative values and a minimum in the water-rich region), features that are slightly but significantly more marked in D2O than in H2O. The V 2 E values in the water-diluted region and at x1, =0 are more negative for D2O than for H2O.  相似文献   

2.
Thermosensitive star-shaped poly(2-isopropyl-2-oxazoline) (molar mass M≈21000 g mol?1) in D2O solution was studied by the static and dynamic light scattering methods. The behavior of the polymer investigated in deuterated water is similar qualitatively to that observed previously in undeuterated water. At the same time, the considerable quantitative changes of polymer behavior in D2O were seen. Deuterium substitution of solvent affects the phase transition temperature by decreasing its value by 1°C. The temperature interval of phase transition in D2O solution expands (by about 1°C) in comparison with that in H2O solution.  相似文献   

3.

The oxidation reaction of phenyl hydrazine (Phh) by hexacyanoferrate ([Fe(CN)6]3?) has been studied in water‐in‐oil (w/o) microemulsion media. The kinetic profile of the reaction was investigated as a function of [Phh], droplet size, and droplet concentration. Comparison of the kinetic profiles of the reaction in microemulsion, water‐urea, and water‐AOT‐urea media indicates that the kinetic profile of the reaction in microemulsion shows a behavior similar to that of the reaction in water‐AOT‐urea medium at 4 M urea. An initial increase and then a decrease in kobs is observed with increasing molar ratio, Wo(=[H2O]/[AOT]) at constant [AOT] (=0.4 M), whereas kobs decreases upon increasing the AOT concentration at constant molar ratio.  相似文献   

4.
Densities of 3-methylpyridine (3-MP) + water and 3-methylpyridine + heavy water were measured in the 3-MP mole fraction range 0.002–0.04 from 298 to 318 K. The excess molar volumes of 3-MP + D2O mixtures were found to be more negative than those of 3-MP + H2O mixtures. The partial molar volume of 3-MP at infinite dilution is smaller in D2O than in H2O which suggests that 3-MP causes a structure-breaking effect in water which is more pronounced in D2O. It was found that the volume change with concentration in dilute solutions of 3-MP in water and heavy water can be adequately described by the pair-wise interaction of the solute molecules. The molal volume second-virial coefficient, V xx , is positive indicating that the water molecules are less structured in the cospheres of the solute pairs than in the bulk solvent. The temperature dependence of V xx displays a maximum at around 308 K in the case of D2O solutions, whereas V xx increases almost linearly with temperature in H2O solutions.  相似文献   

5.
Chand, A., McQuillan, A.R. and Fenby, D.V., 1979. Thermodynamic study of systems with lower critical solution temperatures: H2O + (C2H5)3N, D2O + (C2H5)3N. Fluid Phase Equilibria, 2: 263–274.Molar excess enthalpies and molar excess volumes are reported for the systems H2O + (C2H5)3N and D2O + (C2H5)3N at temperatures below and above their lower critical solution temperatures. The molar excess enthalpies are slightly less exothermic for the D2O system. The molar excess volumes of the H2O and D2O systems are within experimental error of one another. Compositions of conjugate solutions estimated from the calorimetric and volumetric measurements agree with those obtained from published liquid—liquid phase diagrams.  相似文献   

6.
The densities of dilute solutions of H2O and D2O in 1,4-dioxane and tert-BuOD have been measured in the interval 288.15–318.15 K with an error of 2·10–6 g/cm3. The limiting partial molar volumes of D2O and H2O in 1,4-dioxane andtert-butanol have been determined by using an original procedure; the changes in the partial molar volume of water due to H-D substitution in the water molecules have been calculated. The analysis of the temperature dependence of the partial volumes of the components of the binary mixtures H2O (D2O) + 1,4-dioxane and H2O (D2O) +tert-BuOH (tert-BuOD) showed on the basis of Maxwell's crossing equations that the addition of small amounts of water significantly alters the structure of the unary organic solvent. In the presence of trace amounts of water the expansibility of 1,4-dioxane increases and that oftert-butanol decreases.For previous communication, see [1].Institute of the Chemistry of Nonaqueous Solutions, Russian Academy of Sciences, Ivanovo 153018. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 3, pp. 568–571, March, 1992.  相似文献   

7.
Densities of H2O and D2O solutions in dimethylsulfoxide, with solute mole fractions ranging up to 0.037, were measured with an uncertainty of 1.5×10?5?g?cm?3 at eight temperatures between 293.15 and 328.15?K (with a step of 5?K) under atmospheric pressure using a sealed vibrating-tube densimeter. Apparent molar volumes and isobaric expansibilities (down to infinite dilution) of water isotopologues, as well as excess molar volumes of both solutes and solutions as a whole, were calculated. The temperature-dependent behavior of H2O??D2O solute isotope effects on the studied molar volumetric characteristics was interpreted by taking into account the structural and related interaction peculiarities of the dissolving medium in question.  相似文献   

8.
The mixture of polyisoprene with sodium-2-diethylhexyl sulfosuccinate (AOT)/H2O/cyclohexane microemulsion is studied with the photon correlation spectroscopy (PCS) and small angel X-ray scattering (SAXS). The water with AOT induces nano-droplets inside the cyclohexane and addition of concentration and length scale of polyisoprene (PI) can change diffusion of nano-droplets. The collective diffusion coefficient (Dc) of nano-droplets decreased with increase in concentration of PI. From SAXS experiment, a spherical–cylindrical transition of nano-droplets by increase in polyisoprene concentrations observed that it can describe behaviour of diffusion.  相似文献   

9.
Using the data of precision densimetry measurements for diluted solutions of H2O and D2O in 1,4-dioxane (1,4-DX) at 288.15 K-318.15 K we calculated the limiting partial molar volumes of the H/D isotopomers of water in dioxane and the excess molar volumes of the stated systems. The water molecules dispersed in 1,4-DX form complexes H-bonded into associates whose packing coefficient slightly exceeds that of the structural aggregates in liquid H2O and D2O. It is concluded that the structure of 1,4-DX is loosened and concomitantly undergoes volume expansion caused by the water microimpurities. The differentiating temperature effect on the volume solvation effects of H2O and D2O in 1,4-DX has been found.Original Russian Text Copyright © 2004 by. E. V. Ivanov, E. Yu. Lebedeva, and V. K. Abrosimov__________Translated from Zhurnal Strukturnoi Khimii, Vol. 45, No. 5, pp. 852–861, September–October, 2004.  相似文献   

10.
Titania nanoparticles have been produced by the controlled hydrolysis of tetraisopropyltitanate (TPT) in sodium bis(2-ethylhexyl)sulfosuccinate (AOT) reverse micelles. Particle formation and aggregation were investigated by static and dynamic light scattering and the chemical species by vibrational spectroscopy. The kinetics of particle formation and aggregation were controlled by varying [H2O]/[AOT] (w 0), [H2O]/[Ti(IV)] and [AOT]/[Ti(IV)]. Nanoparticles, with diameters<10 nm, could be produced at relatively high Ti(IV) concentrations (up to 0.05 M). These nanoparticles aggregated into sols, with colloid sizes of 20 to 200 nm, eventually forming gelatinous precipitates. Different titania phases were produced, depending on the size of the micellar water pool; small pools (w 0<6) yielded amorphous particles, while larger pools (w 0>10) produced anatase.  相似文献   

11.
The enthalpies of precipitation, of AgCl, Agl, Ag2S and silver tetraphenylborate (AgTPB) in water containing reversed sodium bis(2-ethylhexyl) sulfosuccinate (AOT) micelles as a function of the molar concentration ratio R (R=[water]/[AOT]) at various concentrations of AOT and reagent salts, were measured by a calorimetric technique. The molar enthalpies of precipitation are independent of the surfactant concentration but are dependent on the concentration of the reagent salts and the R value. The molar enthalpies for the same processes in bulk water are not approached even at the highest R value. Effects due to the smallness of the microcrystals and to the interactions between ions and micellar interface are discussed. In part from Doctor of Biology thesis of F. Pinio, University of Palermo, Italy.  相似文献   

12.
In this research, CuO–ZrO2 nanoparticles are synthesized using microreactors made of surfactant/water/cyclohexane microemulsions. The effect of different microemulsion variables on the particle size and its distribution, such as water-to-surfactant molar ratio (W 0) and different surfactants are discussed. Three different surfactant types including cationic (CTAB), anionic (AOT), and nonionic (Brij56) are used. Also a different amount of water to surfactant in nano composite synthesis is used. The powders were characterized by DTA/TG, XRD, SEM, EDS, TEM and BET techniques and their physical properties are compared. The results show a decrease of particles size in presence of cationic surfactant. Narrow particles size distribution of the resultant CuO–ZrO2 nanocomposite in presence of cationic surfactant, anionic and nonionic surfactant is compared. Also for AOT surfactant, by raising water to surfactant molar ratio the particles size is increased and the optimum ratio is H2O: Surfactant = 0.32:0.055, respectively.  相似文献   

13.
The solubilities of glycine, alanine, phenylalanine, and proline in H2O and D2O from 283 K to 335 K were determined. It was found that glycine and alanine are less soluble in heavy water than in light water but proline is more soluble in heavy water over the whole temperature range studied. Phenylalanine is more soluble in H2O than D2O below 310 K but above that temperature heavy water becomes a better solvent. An influence of H/D isotope substitution on the enthalpies of solution is also observed. In the case of glycine and alanine enthalpies of solution in heavy water increase by a small amount and in the same time the solution enthalpy for phenylalanine in D2O increases markedly. No change in the solution enthalpy for proline was observed. The isotope effects on solubility and the solution enthalpy are qualitatively discussed.  相似文献   

14.
Densities of solutions of H2O and D2O in nitromethane, with the solute mole fractions ranging up to 0.03, were measured with an error of 1.5 · 10−5 cm3 · mol−1 at (278.15, 288.15, 298.15, 308.15, and 318.15) K using a vibrating-tube densimeter. Apparent and partial volumes and isobaric expansibilities (down to the infinite dilution) of water isotopologues were calculated. The temperature-dependent behavior of D2O–H2O solute isotope effects on the molar quantities studied were described taking into account the structure- and interaction-related peculiarities of the dissolving medium in question.  相似文献   

15.
Densities of H/D-isotopomers mixtures of water (H2O, D2O) and methanol (CH3OH, CD3OH, CH3OD, and CD3OD) over the full range of compositions were measured at 5, 15, 25, 35, and 45°C. Results have been used to calculate molar volumes, excess molar volumes, apparent molar volumes, and isotope effects of the mixtures. The volumetric properties are discussed in terms of the structural changes in water-methanol solutions under the influence of isotope substitution.  相似文献   

16.
Densities of dilute solutions of 1,3-dimethyl-2-imiazolidinone in H2O and D2O, with the solute mole-fractions ranging up to 0.01, have been measured with an error of 1.5 · 10−5 g · cm−3 at (278.15, 288.15, 298.15, 308.15, 313.15, and 318.15) K and atmospheric pressure using a vibrating-tube densimeter. The partial molar volumes of the dissolved DMI (down to the infinite dilution) and solvent (H2O or D2O) as well as the excess molar volumes of the isotopically distinguishable solutions have been calculated. The effects of the solvent isotope substitution, solute concentration and temperature on the volume changes caused by DMI hydration have been considered. The obvious relationship between the D2O–H2O solvent isotope effects on the partial molar volume and enthalpy of solution of DMI has been discovered.  相似文献   

17.
IR relative integrated intensities and half-widths of rocking (R) and wagging (W) bands of water in MnCl2 · 2H2O and CoCl2 · 2H2O are presented at 300 K and 120 K. Departure of observed intensity into DW/DR from those predicted by the fixed dipole model is attributed to anisotropic dynamic changes in dipole during these oscillations. A quantity representing the variation of this anisotropy between W and R oscillations is computed and its origin is discussed. An increase by 20% to 50% in both DW and DR on lowering the temperature has also been discussed.  相似文献   

18.
The kinetics of the formation of hydrogen peroxide by the sonolysis of light and heavy water in argon and oxygen atmospheres was investigated. The sonochemical reaction has a zero order with respect to hydrogen peroxide (H2O2, D2O2, or DHO2). The measurement of the kinetic isotope effect (α), defined as the ratio of the reaction rates in H2O and D2O, carried out under argon gave a value of 2.2±0.3. The observed isotope effect decreases with an increase in the concentration of light water in H2O−D2O mixtures. No isotope effect is displayed in the oxygen atmosphere (α=1.05±0.10). The isotope effect is determined presumably by the mechanism of sonochemical decomposition of water molecules, which includes the H2O−Ar* and D2O−Ar* energy exchange (where Ar* are argon atoms in the3P2.0 excited state) in the nonequilibrium plasma generated by a shock wave, arising upon a cavitation collapse. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 645–649, April, 2000.  相似文献   

19.
Excess volume, partial molar volumes, viscosity, and ultrasound velocity in H2O and D2O solutions of formamide (FM), acetamide (AM), dimethylformamide (DMF), and dimethylacetamide (DMA) were studied at 20°C. The change in the excess volume of the various amide solutions on substitution of D2O for H2O varies both in magnitude and direction. On the other hand, the isotope effect on the ultrasound velocity behaves similarly in all cases. Its magnitude is greatest for pure water and then decreases monotonically with increasing amide concentration. Although the behavior of the concentration dependence of viscosity is similar to that for ultrasound velocity, the isotope effect on the viscosity behaves in a different way. For methyl-substitued amides, a maximum isotope effect is observed at amide mole fraction 0.2, but the isotope effect for FM and AM increases monotonically with increasing amide concentration. The differences in the behavior of amides in aqueous solution are discussed in terms of their interactions with water.  相似文献   

20.
The photophysical parameters of two probes with largely different hydrophobic character, namely, coumarin 1 and coumarin 343, are investigated in sodium bis‐(2‐ethylhexyl)sulfosuccinate (AOT)/hexane/water reverse micelles at various water/AOT molar ratio w0. Correlation of photophysical parameters such as fluorescence quantum yield, fluorescence lifetime, and emission maxima with w0 indicate distinctly different trends below and above w0≈7 for both probes. The variation of the average rotational correlation times obtained from fluorescence anisotropy decays for both probes in reverse micelles further corroborate the above observation. Similar studies were also performed in nonaqueous reverse micelles with acetonitrile as polar solvent. Similar to aqueous reverse micelles, breaks in the photophysical parameters with increasing acetonitrile/AOT molar ratios w0 were also observed in these cases, although at a much lower w0 value of 3. The present results indicate that around w0≈7 for aqueous reverse micelles (and around w0≈3 for nonaqueous reverse micelles) a distinct change occurs in the probe microenvironment, which is rationalized on the basis of the relative populations of interfacial and core water. We propose that until the ionic head groups and counterions are fully solvated by polar solvents, that is, up to w0≈7 (or w0≈3), the interfacial water population dominates. Above these molar ratios coalescence of excess water molecules with each other to form truncated H‐bonded water clusters leads to a sizable population of core water. This is further substantiated by changes in the IR absorption spectra for the O? D stretching mode of diluted D2O in reverse micelles with varying w0. Critical comparison of the present results with relevant literature reports provide clear support for the proposals made on water structure in reverse micelles. The role of relative size of the probe and the reverse micelles for differences in polar solvent to AOT ratios (w0=7 and w0=3) in the observed breaks in the two types of reverse micelles is also discussed.  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