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.
The density, isothermal compressibility, and thermal expansion coefficient of macroscopically single-phase n-heptane–water–sodium dodecyl sulfate–n-pentanol microemulsions are measured by a precision dilatometric method at 25°C. The internal pressure and molar volume of the investigated microemulsion systems are calculated. It was shown that the bulk properties of direct microemulsions with various contents of the dispersed phase markedly differ.  相似文献   

2.
In this work we report about a new rare-earth oxoborate β-Dy2B4O9 synthesized under high-pressure/high-temperature conditions from Dy2O3 and boron oxide B2O3 in a B2O3/Na2O2 flux with a walker-type multianvil apparatus at 8 GPa and 1000°C. Single crystal X-ray structure determination of β-Dy2B4O9 revealed: , a=616.2(1) pm, b=642.8(1) pm, c=748.5(1) pm, α=102.54(1)°, β=97.08(1)°, γ=102.45(1)°, Z=2, R1=0.0151, wR2=0.0475 (all data). The compound exhibits a new structure type which is built up from bands of linked BO3- (Δ) and tetrahedral BO4-groups (□). The Dy3+-cations are positioned in the voids between the bands. According to the conception of fundamental building blocks β-Dy2B4O9 can be classified with the notation 2Δ6□:Δ3□=4□=3□Δ. Furthermore we report about temperature-resolved in situ powder diffraction measurements and IR-spectroscopic investigations on β-Dy2B4O9.  相似文献   

3.
Interactions between α-CD and three alkyl trimethyl ammonium bromides, a homologues series of surfactants, in aqueous solutions have been investigated with titration microcalorimetry at 298.15 K. The results are discussed in the light of the amphiphilic interaction and the iceberg structure of water molecules existing around the hydrophobic tail of the surfactant. The stoichiometry of the host–guest complex changes from 1:1 to 2:1, as the number of carbon atoms (n) in the hydrophobic chain, CnH2n+1, increases from 8 to 14. All the complexes are quite stable, with the apparent experiential stability constants being β1 = 2.65 × 103 dm3-mol−1, β2 = 4.85 × 106 dm6-mol−2, β2 = 6.50 × 106 dm6-mol−2, respectively for n = 8, 12, 14. All the complexation processes are shown to be enthalpy driven, and the standard enthalpy effect (−ΔH0) increases while standard entropy change (ΔS0) decreases with elongation of the hydrophobic chain.  相似文献   

4.
Dichlorocobalt(III) complexes of (2S,5S,9S)-trimethyltriethylenetetraamine (L1) and (2S,5R,9S)-trimethyltriethylenetetraamine (L2) have been prepared. Both L1 and L2 coordinate to the cobalt(III) ion to give three isomers: Λ-cis-α, Δ-cis-β, trans isomers for L1 and Δ-cis-α, Δ-cis-β, trans isomers for L2. Each of the trans-dichloro complexes of the two ligands have been isomerized stereospecifically to the cis-α-dichloro complex in methanol, and each of the cis-α-dichloro complexes stereospecifically to the trans-diaqua complex in water. Both the geometrical and optical inversions took place at the same time in the observed stereospecific isomerizations.  相似文献   

5.
6.
The effects of the variables of head group structure and salt concentration on microemulsions formed in mixtures of water, alkyl ethylene glycol ethers (CkOC2OCk), andn-alkyl β- -glucopyranosides (CmβG1) are explored. Phase behavior of mixtures containing an anomer of the surfactant (n-alkyl α- -glucopyranoside, CmαG1), or surfactants with long head groups (n-alkyl maltopyranosides, CmG2), or NaCl or NaClO4as electrolyte are systematically reported as a function of temperature and composition. The substitution ofn-alkyl α- -glucopyranosides forn-alkyl β- -glucopyranosides causes precipitation under some conditions in all mixtures studied. These solubility boundaries begin in the water–surfactant binary mixture at the Krafft boundary, then extend to high concentrations of both surfactant and oil. Increasing the effective length of the surfactant head group by adding CmG2to water–CkOC2OCk–CmβG1mixtures moves the phase behavior dramatically up in temperature when even small amounts of CmG2are used. Adding a lyotropic electrolyte, NaCl, to water–CkOC2OCk–CmβG1mixtures moves the phase behavior down in temperature, while the hydrotropic electrolyte NaClO4moves the phase behavior up in temperature.  相似文献   

7.
When the reaction conditions are deliberately controlled by the pH, two different polynuclear manganese complexes, (Δ, Λ)-{Mn3(phen)2 (CH3COO)6} (1) and [Mn(phen)Cl2]n (2) (phen = 1,10-phenanthroline), have been synthesized from the same raw materials. The structural analyses show that 1 has a structure formed by neutral chiral linear trinuclear molecules, while 2 has a structure consisting of one-dimensional infinite chains. A study of the temperature dependent magnetic susceptibilities reveal that 1 is an antiferromagnetically coupled trimer molecule while 2 shows ferromagnetic interactions within the chain.  相似文献   

8.
The dielectric properties of mixed monolayers of per-(6-amino-2,3-di-O-hexyl) β-CD hydrochloride (NH3-β-CD-OC6) and 1,2 dipalmitoyl, 3-sn-phosphatidic acid (DPPA) have been assessed using surface potential measurements at constant area. From the comparison of these surface potential (ΔV) versus surface density (δ) relationships with those of surface pressure (π) against surface density (δ) it was apparent that the increase in the NH3-β-CD-OC6 content in mixed films gave rise to a gradual increase in the saturation value of the surface potential (ΔVmax). This potential for pure DPPA was found to be equal to 396 mV and for pure CD 554 mV. The ΔVmaxvalues reflect the onset of reorientation effects that arrive at molar areas before the collapse of these films. Independently of reorientation effects, the obtained results strongly indicate that the dipolar term contributing to the overall ΔVvalue was for NH3-β-CD-OC6 due to the hydration of its NH+3group. For both DPPA and NH3-β-CD-OC6 molecules the contribution of the electric double layer (Ψ) was calculated and was found for DPPA and NH3-β-CD-OC6 to be equal to −249 and +252 mV, respectively. These calculated Ψ values made possible the evaluation of dipole moments for NH3-β-CD-OC6 and DPPA monolayers which revealed a marked difference in dipolar properties between these two film forming components. In contrast to DPPA which exhibited a decrease in the surface dipole moment (μ) with the decrease inA, NH3-β-CD-OC6 displayed an increase in μwith the decrease inAforAvalues above 580 Å2. Below this value μdecreases with decreasing molecular area and this variation arises from a change in the polarity of the electric double layer arising from interactions with the complementary anion. The differences in dielectric properties between the two film forming molecules have been attributed to modification, during compression, in the structure of the interfacial water bound to the cyclodextrin.  相似文献   

9.
A temperature study was performed on micelle formation of a series of homologous cationic surfactants having organic counterions (alkanesulfonates) with carbon numbers ranging from 1 to 4: dodecylammonium salts of methanesulfonate (DAMS), ethanesulfonate (DAES), propanesulfonate (DAPS), and butanesulfonate (DABS) in water. The critical micelle concentrations (CMCs) and the degree of counterion binding (β) were determined at different temperatures ranging from 5 to 50°C by means of conventional electric conductance measurements. From the temperature dependence of β as well as CMC, Gibbs energy ΔG0m, enthalpy ΔH0m, and entropy ΔS0m, on micelle formation, were estimated for the respective surfactants. As for the temperature dependence of CMC for these surfactants, the temperature-CMC curves have a minimum around 30°C and show that the CMC at each temperature is lowered by about 3 mmol dm-3 per methylene group in the alkyl chain of the counterions. The relationship between β and temperature suggested that the counterion of MS- behaves most similarly to common univalent ions such as halide ions. In contrast, PS- and BS-, having a stronger ability to lower CMC and to promote association of surfactant ions with counterions as well as of surfactant ions themselves, behave more like those of surfactant ions, and ES- shows the most complicated character between those of common univalent ions and organic ions. However, the temperature dependence of enthalpy change, ΔH0m demonstrates that these four surfactants are divided into two groups: (1) DAMS and DAES and (2) DAPS and DABS. In addition, the entropy change ΔS0m as a function of alkyl chain length gives evidence that the contribution of the entropy term to the Gibbs energy on micelle formation clearly separates between DAES (m = 2) and DAPS (m = 3). A similar discontinuity is found even in the plot of ΔG0m versus carbon atom number of alkyl chain, m, and in the plot of ΔG0m versus estimated hydrodynamic radius of counterions. All the results obtained have indicated that lengthening the alkyl chains initially hinders micelle formation, but the longer chains are markedly effective in lowering the CMC and probably in increasing the aggregation number, owing to enhanced hydrophobic interaction between counterion and the micellar surface and/or core.  相似文献   

10.
Chloride complexation of cobalt(II), nickel(II) and zinc(II) ions has been studied by calorimetry and spectrophotometry in N-methylformamide (NMF) containing 1.0 mol-dm− 3 (n-C4H9)4NClO4 as an ionic medium at 298 K. A series of mononuclear complexes, MCln(2 -n) + (M=Co, Ni and Zn) with n = 1, 3 and 4 for cobalt(II), n = 1 for nickel(II), and n = 1–4 for zinc(II), are formed and their formation constants, enthalpies and entropies were obtained. It revealed that complexation is suppressed significantly in NMF relative to that in N,N-dimethylformamide (DMF) in all metal systems examined. The suppressed complexation in NMF is mainly ascribed to the smaller formation entropies in NMF reflecting that the solvent–solvent interaction or solvent structure in the bulk NMF is much stronger than that in the bulk DMF. Formation entropies, Δ S1o, of the monochloro complex in DMF, dimethyl sulfoxide and NMF are well correlated with the Marcus’ solvent parameter, Δ Δv So/R, according to Δ S1o/R = aΔ Δv So/R+b. The a value is negative and similar in all metal systems examined, whereas the b value depends on the metal system. When a gaseous ion is introduced into a solvent, the ionic process of solvation is divided into two stages: the ion destroys the bulk solvent structure to isolate solvent molecules at the first stage and the ion then coordinates a part of isolated solvent molecules around it at the second stage. We propose that the a and b values may reflect the changes in the freedom of motion of solvent molecules at the first and second stages, respectively, of the ionic process of solvation.  相似文献   

11.
The Conder and Young (CY) and the peak maximum (PM) methods were used to estimate the retention time of n-alkane probes on chemithermomechanical pulp (CTMP) wood fibers treated with a low molecular weight grade phenol-formaldehyde resin (PFR). Thermodynamic functions (ΔHao, ΔGao, and ΔSao) and the London dispersive component of the surface energy were derived from these retention times. Treated wood fibers show a high energy surface due to the presence of the thermoset resin on their surface. Values of ΔHao obtained from the CY method were higher than those obtained with the PM method at relatively high temperatures and with relatively low molecular weight alkanes. The results from the two methods were identical at low temperature (293 K) and with the relatively high molecular weight alkane n-undecane.  相似文献   

12.
Okada and Matsuura's transport equations for pervaporation give rise to three fundamental parameters, namely, interfacial saturation vapor pressure P*, liquid transport parameter A/δ, and vapor transport parameter B/δ. The effects of the chemical nature of the membrane material and the upstream operating pressures of 101.3 and 303.9 kPa on the above parameters were investigated from the pervaporation data at laboratory temperature (24°C) for water and ethanol using a cellulose acetate butyrate membrane. The results show that the P*, values are essentially unaffected by the upstream pressure, and that they are generally higher than the literature values of saturation vapor pressure at 24°C. Further, the values for A/δ and B/δ tend to increase with increased upstream pressure for both systems studied. These results are discussed.  相似文献   

13.
The pressure induced structural transition of NaBH4 from β-NaBH4 (tetragonal-P421c) to γ-NaBH4 (orthorhombic-Pnma) is investigated by ab initio plane-wave pseudopotential density functional theory method (DFT). The BaSO4-type structure of orthorhombic high-pressure phase is testified theoretically for the first time. The calculated transition pressure of β-NaBH4 (tetragonal-P421c) to γ-NaBH4 (orthorhombic-Pnma) is 9.66 GPa and the orthorhombic high-pressure phase is stable up to 30 GPa. Our results agree well with previous experimental results and demonstrate that high-pressure phase transition from β-NaBH4 to γ-NaBH4 may occur at low temperature. At last, the pressure effects on the electronic structures of α-, β- and γ-NaBH4 are discussed.  相似文献   

14.
The mutual solubility of components and critical solution points in the ternary system water-isopropyl alcohol-n-dodecane were studied by a visual-polythermal technique at saturated vapor pressure in the temperature range 5–120 °C. The mixture composition in the critical solution point was studied in relation to temperature.__________Translated from Zhurnal Prikladnoi Khimii, Vol. 78, No. 3, 2005, pp. 394–397Original Russian Text Copyright © 2005 by Sinegubova, Il’in, Cherkasov.  相似文献   

15.
The values of density (ρ), viscosity (η) and speed of sound (u) have been measured for binary liquid mixtures of γ-butyrolactone (GBL), δ-valerolactone (DVL), and ε-caprolactone (ECL) with N-methylacetamide (NMA) over the whole composition range at T = (303.15 to 318.15) K and atmospheric pressure. From these data, excess molar volume (VE), deviation in viscosity (Δη), and deviation in isentropic compressibility (Δκs), are calculated. The results are fitted to a Redlich–Kister type polynomial equation to derive binary coefficients and standard deviations.  相似文献   

16.
25,26,27,28-Tetrahydroxycalix[4]arene-5,11,17,23-tetrasulphonate was in a heat-conduction calorimeter titrated with α,ω-alkyl diammonium ions, n=(3–7), in an aqueous solution at pH 7.1. Apparent concentration binding constants (Kc) and enthalpy changes are reported for a 1 : 1 binding model at minimized concentration ranges. The standard Gibbs energy change, ΔG0=−RTln Kc, has a non-linear dependence of the alkyl chain length, n. The interaction is strongest at n≥5. A similar trend is observed for the enthalpy change. However, the fit of the experimental data to the 1 : 1 model was quite poor and neither 1 : 2 nor 2 : 1 models gave any further improvement. Concentration-dependent results were also obtained. It is therefore concluded that the interaction observed is more complex.  相似文献   

17.
The sitting-atop complexation of meso-tetraarylporphyrins and its para-substituted derivatives (H2t(4-X)pp, X:H, Br, Cl, CH(CH3)2, OCH3, CH3), as electron donors, with zirconyl, as an electron acceptor, have been investigated spectrophotometrically in chloroform. The mole ratio studies based on physicochemical techniques were employed clearly and revealed the formation of 1:1 sitting-atop complexes which was confirmed by UV–vis, 1H NMR and IR spectroscopic data. The value of the formation constant was estimated for each complex using a nonlinear optimization of the complex absorbance vs. mole ratio data by package KINFIT. The results showed that the stability of these complexes decreases with the temperature enhancement. Thermodynamic parameters, ΔG°, ΔH° and ΔS°, of the SAT complexes have been determined from the temperature dependence of formation constants by Van’t Hoff equation. Also, the influence of the substituents of the aryl rings in H2t(4-X)pp on the stability of the SAT complexes is discussed.  相似文献   

18.
REDOR technique was applied to natural abundance 13C nuclei coupled to a singly labeled 15N nucleus to determine the 13C, 15N interatomic distances simultaneously in crystalline ammonium [15N] -glutamate monohydrate (1). Consequently, the interatomic C–N distances between 15N and 13C=O, 13Cα, 13Cβ, 13Cγ, and 13Cδ carbon nuclei for 1 were determined with a precision of ±0.15 Å, after the experimental conditions such as the location of samples in the rotor, length of π pulse etc. were carefully optimized. 13C-REDOR factors for three spin system, (ΔS/S0)CN1N2, and the sum of two isolated 2-spin system, (ΔS/S0)*=(ΔS/S0)CN1+(ΔS/S0)CN2, were further evaluated by the REDOR measurements on isotopically diluted 1 in a controlled manner. Subsequently, the intra- and intermolecular C–N distances were separated by searching the minima in the contour map of root mean square deviation (RMSD) between the theoretically and experimentally obtained (ΔS/S0)* values against two interatomic distances, rC–N1 and rC–N2. When the intramolecular C–N distance (rC–N1) of the particular carbon nucleus is substantially shorter than the intermolecular one (rC–N2), C–N distances within a single molecule were obtained with an accuracy of ±0.06 Å as in the cases of C=O, Cα and Cβ carbon nuclei. C–N distances between the molecule in question and the nearest neighboring molecules can be also obtained, although accuracy was lower. On the contrary, it was difficult to determine the interatomic distances in the same molecule when the intermolecular dipolar contribution is larger than the intramolecular one as in the case of Cδ carbon nucleus.  相似文献   

19.
The standard molal potential differences (Em∘) have been determined for the cell: CdHgx(two phase) | CdCl2(m), H2O(1 − w), 2-butanol (w) | AgCl(s) | Ag(s) in aqueous mixtures of low mass fraction of 2-butanol (w2-butanol = 0.05, 0.10, and 0.15) by using the literature data for the stability constants of the chlorocadmium complexes and the present potentiometric data for this cell at five temperatures from (293.15 to 313.15) K and at 10 molalities of CdCl2 from (0.002 to 0.02) mol-kg−1. The resulting values of Em have been used to calculate the standard thermodynamic quantities (ΔrG, ΔrH, and ΔrS) for the cell reaction, the stoichiometric mean molal activity coefficients (γ±) of CdCl2, and the standard thermodynamic functions for CdCl2 transfer (Δt G∘, Δt H∘, and Δt S∘) from water to the examined aqueous mixtures of 2-butanol. The values obtained have been compared with the analogous literature data for aqueous mixtures of 2-butanone; standard thermodynamic quantities for transfer of CdCl2 and HBr from water to mixtures containing the same mass fraction of 2-butanol have also been compared. For both electrolytes, these quantities show analogous trends with the alcohol content. This transfer process is nonspontaneous and endothermic. Enthalpy and entropy are evidently influenced by structural changes.  相似文献   

20.
Titanium(IV) chloride reacts with free base meso-tetraarylporphyrin and its ortho, meta and para-substituted derivatives (H2T(X)PP; X: OCH3, CH3 and Cl) for formation of sitting-atop (SAT) complexes, [TiCl4(H2T(X)PP)]. The computer fitting of the variation of the absorbance versus mole ratio by KINFIT program was used for calculation of the formation constants of these complexes in chloroform. Thermodynamic parameters, ΔG°, ΔH°, ΔS°, have been determined and the influence of the temperature and the substituted aryl groups (electronic and steric effects) in the free base porphyrins on the stability of the SAT complexes was studied.  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