共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
Y. B. Tewari D. E. Martire S. P. Wasik M. M. Miller 《Journal of solution chemistry》1982,11(6):435-445
From thermodynamics and certain assumptions it is shown that, under the usual experimental conditions, the octanol-water partition coefficient (Ko/w) of a given organic liquid should be the same whether the substance is partitioned neat or as part of a mixture. Measurements of several mixtures of n-propylbenzene (log Ko/w=3.71±0.04)+ethylbenzene (log Ko/w=3.16±0.01) clearly confirm this. It is also shown that the aqueous solubility (Sw) of a neat organic liquid can be related to its aqueous solubility (S
w
), when it is present at volume fraction in an organic liquid mixture, by S
w
=Sw, where is its activity coefficient in the organic mixture. The measured S
w
values for n-propylbenzene + ethylbenzene (1), n-hexane + nitrobenzene (>1) and di-isopropyl ether + chloroform (<1) are found to be in good agreement with the predicted values (average differences of, respectively, <2%, 8% and 6%). In general, the bounds on S
w
are expected to be 0w
w. 相似文献
3.
《中国化学会会志》2018,65(2):217-224
A new and simple potentiometric method is developed to determine the stability constants of cysteine complexes with Mn(II), Cu(II), Zn(II), Cd(II), and Ce(III) metal ions using a modified Bjerrum method. Potentiometric titrations were performed in water and water/dioxane mixtures at five different temperatures. The least and highest stable complexes were those of Cu(II) and Zn(II), respectively. Furthermore, by conducting the experiments at temperatures of 15, 20, 25, 35, and 45 °C, the thermodynamic parameters ΔH°, ΔS°, and ΔG° for the pertinent complexes were calculated. The results show that the ΔH° values of the complexes are positive except for Zn(II). Negative values of ΔG° are evidence for the spontaneity of the reactions. 相似文献
4.
5.
A conventional solid-phase microextraction (SPME) method combined with liquid–liquid extraction was applied under equilibrium and nonequilibrium conditions to determine the partition coefficients (Kdoc) of 25 polychlorinated biphenyl congeners (PCBs) between Sigma–Aldrich humic acid (HA) and water. The values of log Kdoc determined with equilibrium SPME were linearly correlated with the logarithm of the octanol–water partition coefficients (Kow) for PCB congeners at log Kow < ∼7.2, but the trends were disrupted for log Kow from ∼7.2 to 8.18. In addition, short-term (5 min to 4 days) and long-term (5–44 days) uptake profiles of PCBs were established, from which a pseudo-equilibrium for sorption of PCBs was revealed at ∼4 days of extraction. To understand this phenomenon, the uptake profiles were fitted with two equations (one equation is often used for pure water samples and the other one is applicable for samples containing complex matrices) derived from a first-order kinetics model. Subsequently, Kdoc values obtained through kinetic approaches were compared with those acquired from equilibrium SPME. The comparison of Kdoc values indicated that the pseudo-equilibrium was caused by the slow desorption of PCBs from HA rather than the biphasic desorption mechanism. 相似文献
6.
Rasmus LundsgaardGeorgios M. Kontogeorgis Ioannis G. Economou 《Fluid Phase Equilibria》2011,306(2):162-170
Accurate partition coefficient data of migrants between a polymer and a solvent are of paramount importance for estimating the migration of the migrant over time, including the concentration of the migrant at infinite time in the two solvents. In this article it is shown how this partition coefficient can be estimated for both a small hydrophilic and a hydrophobic organic molecules between squalane (used here to mimic low density poly ethylene) and water/ethanol solutes using thermodynamic integration to calculate the free energy of solvation. Molecular dynamics simulations are performed, using the GROMACS software, by slowly decoupling of firstly the electrostatic and then the Lennard-Jones interactions between molecules in the simulation box. These calculations depend very much on the choice of force field. Two force fields have been tested in this work, the TraPPE-UA (united-atom) and the OPLS-AA (all-atom). The computational cheaper TraPPE-UA force field showed to be more accurate over the whole range of systems compared to the OPLS-AA force field. Moreover, some of the calculations were done with five different water models to investigate the influence of the specific water model on the calculations. It was found that the combination of the TraPPE-UA force field and the TIP4p water model gave the best results. Based on the methodology proposed in this article, it is possible to obtain good partition coefficients only knowing the chemical structure of the molecules in the system. 相似文献
7.
8.
9.
?ukasz Gurzyński Mariusz Makowski Lech Chmurzyński 《The Journal of chemical thermodynamics》2006,38(12):1584-1591
By using the potentiometric method, acidity constants have been determined in systems of tri- and tetra-substituted pyridine N-oxides. The potentiometric measurements in systems of four 4-chloropyridine N-oxide derivatives containing the chlorine atom at position 4 to the NO2 group and four bromine counterparts were carried out in polar non-aqueous solvents, viz. amphiprotic methanol (MeOH) and aprotic protophilic dimethyl sulfoxide (DMSO). It was found that in all the systems studied the pKa values were readily determinable (as indicated by small standard deviations) in MeOH, whereas in DMSO large standard deviations were obtained making the pKa values either hardly determinable or indeterminable from potentiometric measurements. Furthermore, it was demonstrated that the acidity constants of protonated N-oxides studied in MeOH changed according to the sequence of their acidity constants in water. It was also found that in the polar solvents studied, i.e. in the amphiprotic methanol and the highly basic aprotic dimethyl sulfoxide, the cationic homo-conjugation equlibrium constants could not be determined using potentiometric method. Also, by using ab initio methods at the RHF and MP2 levels and the PCM model, utilizing the Gaussian 6-31++G∗∗ basis set, energies and Gibbs free energies of the protonation reactions of the N-oxides have been determined. The energy parameters have been compared with acidity constants of the protonated N-oxides determined by potentiometric titration in methanol to establish a correlation between these approaches. 相似文献
10.
Raquel Ares Margarita López-Torres Nina Gómez-Blanco Jesús J. Fernández 《Journal of organometallic chemistry》2008,693(24):3655-3667
Treatment of the chloro-bridged dinuclear complex [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}(μ-Cl)]2 (1) with homobidentate [P,P], [As,As], [N,N], and heterobidentate [P,As], [P,N] ligands in a 1:1 molar ratio gave the dinuclear complexes [{Pd[3,4-(MeO)2C6H2C(H)N(Cy)-C6,N](Cl)}2{μ-L}] (L = Ph2PC4H6(NH)CH2PPh2 (2); Ph2As(CH2)2AsPh2 (3); 1,3-(NH2CH2)2C6H4 (4); Ph2P(CH2)2AsPh2 (5); Ph2P(CH2)2NH2 (6)), with the bidentate ligands bridging the two cyclometallated fragments.The reaction with the homobidentate ligands in a 1:2 molar ratio in the presence of NaClO4 afforded the mononuclear compounds [[Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{L-P,P}][ClO4] (L = Ph2PC4H6(NH)CH2PPh2 (7); (o-Tol)2P(CH2)2P(o-Tol)2 (8)), [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{Ph2As(CH2)2AsPh2-As,As}][ClO4] (9) and [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{L-N,N}][ClO4] (L = NH2(CH2)3NH2 (10); NH2(C6H8)CH2(C6H8)NH2 (11); 1,3-(NH2CH2)2C6H4 (12); 1,3-(NH2)2C5H3N (13); NH2(C6H4)O(C6H4)NH2 (14); NMe2(CH2)2NMe2 (15)), in which the chloro ligands are absent and the bidentate ligands are chelated to the palladium atom.Reaction of 1 with Ph2P(CH2)2AsPh2 in 1:2 molar ratio in acetone in the presence of NH4PF6 afforded the analogous mononuclear compound [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{Ph2P(CH2)2AsPh2-P,As}][PF6] (16); whereas reaction with Ph2P(CH2)3NH2 gave [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{Ph2P(CH2)3N(CMe2)-P,N}][PF6] (17), derived from intermolecular condensation between the aminophosphine and acetone. Condensation of the NH2 group was precluded by change of solvent, using dichloromethane.Iminophoshines also reacted with 1 in 1:2 molar ratio in acetone to give a new series of mononuclear cyclometallated complexes: [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{L-P,N}][ClO4] (L = Ph2PC6H4C(H)NCy (20); Ph2PC6H4C(H)NC(CH3)3 (21); Ph2PC6H4C(H)NNMe2 (22); Ph2PC6H4C(H)NNHMe (23); Ph2PC6H4C(H)NNHPh (24)). Analogous complexes with a stable P,O-chelate were obtained using bidentate [P,O] donor ligands: [Pd{3,4-(MeO)2C6H2C(H)N(Cy)-C6,N}{L-P,O}][Cl] (L = 2-(Ph2P)C6H4CHO (25); Ph2PN(Me)C(O)Me (26)).The crystal structures of compounds 1, 5, 15, 16, 18, 20 have been determined by X-ray crystallography. 相似文献
11.
Electrochemical measurements are done on (water + NaBr + K3PO4 + glycine) mixtures at T (298.15 and 308.15) K by using (Na+ glass) and (Br− solid-state) ion selective electrodes. The mean ionic activity coefficients of NaBr are determined at five NaBr molalities (0.1, 0.3, 0.5, 0.7, and 1) in the above mixtures. The activity coefficients of glycine are evaluated from mean ionic activity coefficients of sodium bromide. The ratio of mean ionic activity coefficient of NaBr in the (water + NaBr + K3PO4 + glycine) mixtures to the mean ionic activity coefficients of NaBr at the same molalities in the (H2O + NaBr) mixtures are correlated by using a new expression. 相似文献
12.
13.
Henry's constants of n-alkanols (methanol to n-hexanol) in water were measured at temperatures between 40°C and 90°C using a recently developed headspace gas chromatographic technique. The data were in good agreement with literature data when available. The consistency of the data was verified by comparing calculated partial molar enthalpies with calorimetric values. The temperature dependence of dimensionless Henry's constants was fitted with the classical van't Hoff equation and an empirical correlation was established for the dimensionless Henry's constants as a function of temperature and number of carbon atoms in the n-alkanol. 相似文献
14.
Fanny d'Orlyé 《Tetrahedron》2005,61(41):9670-9678
In DMF at 80 °C, a Pd0 complex is generated in situ from the dimeric P,C-palladacycle (1) in the absence of any reducing agents, presumably via a reductive elimination. The Pd0 complex formed in an endergonic equilibrium has been trapped and stabilized by an additional P(o-Tol)3 and has been detected in cyclic voltammetry by its oxidation peak. Its formation is favored by acetate anions (often used as base in Heck reactions) via the formation of a monomeric anionic P,C-palladacycle ligated by acetate ions. As postulated, P,C-palladacycles are a reservoir of monophosphine-Pd0 complexes active in oxidative additions with aryl halides. 相似文献
15.
Chao Liang Shu‐ying Han Jun‐qin Qiao Hong‐zhen Lian Xin Ge 《Journal of separation science》2014,37(22):3226-3234
A strategy to utilize neutral model compounds for lipophilicity measurement of ionizable basic compounds by reversed‐phase high‐performance liquid chromatography is proposed in this paper. The applicability of the novel protocol was justified by theoretical derivation. Meanwhile, the linear relationships between logarithm of apparent n‐octanol/water partition coefficients (logKow′′) and logarithm of retention factors corresponding to the 100% aqueous fraction of mobile phase (logkw) were established for a basic training set, a neutral training set and a mixed training set of these two. As proved in theory, the good linearity and external validation results indicated that the logKow′′–logkw relationships obtained from a neutral model training set were always reliable regardless of mobile phase pH. Afterwards, the above relationships were adopted to determine the logKow of harmaline, a weakly dissociable alkaloid. As far as we know, this is the first report on experimental logKow data for harmaline (logKow = 2.28 ± 0.08). Introducing neutral compounds into a basic model training set or using neutral model compounds alone is recommended to measure the lipophilicity of weakly ionizable basic compounds especially those with high hydrophobicity for the advantages of more suitable model compound choices and convenient mobile phase pH control. 相似文献
16.
Yashar M. Naziev 《The Journal of chemical thermodynamics》2005,37(12):1268-1275
The (p, ρ, T) of methanol, ethylbenzene and (methanol + benzene) and (methanol + ethylbenzene) at temperatures between (290 and 500) K and pressures in the range (0.1 to 60) MPa have been measured with a magnetic suspension densimeter with an uncertainty of ±0.1%. Our measurements with methanol deviate from the literature values by less than 0.2%. The (p, ρ, T) measurements were fitted with experimental uncertainties by an empirical equation. The temperature and mole fraction dependence of the coefficients of the equation of state are presented. 相似文献
17.
Stoichiometry and equilibrium study of copper‐ligands including mercaptobenzoxazole (MBO), 4‐propyl 2‐thiouracyl (PTU), methyl‐2‐pyridylketone oxime (MPKO), phenyl‐2‐pyridylketone oxime (PPKO), 4,6‐dihydroxy‐2‐mercaptopyrimidine (DHMP), N,N′‐phenylene bis(salicylaldimine) (PBS) and 1,2‐bis(2‐hydroxyphenyl)naphtaldiimine (BHNPDI) were conducted in aqueous and nonaqueous solution by potentiometry and spectrophotometry. Stability constants of the complexes are determined at 25 ± 1 °C and 0.1 or 0.05 M ionic strength in water or acetonitrile solvents. Oximes ligand protonation constants and copper‐ligands complexes' stability and hydrolysis constants were calculated using the BEST program in aqueous solution. The stability constants of copper‐ligands complexes were calculated using the KINFIT program in acetonitrile solution. The results of these two methods are made self‐consistent, then rationalized assuming an equilibrium model including the species, ML, MLH, MLOH and ML2 (where the charges of the species have been ignored for the sake of simplicity) (L = MBO, PTU, MPKO, PPKO, DHMP, BHNPDI and PBS). 相似文献
18.
Wen Zhang 《Tetrahedron letters》2004,45(48):8921-8924
Axially dissymmetric P,S-heterodonor ligand L3 synthesized from BINOL is an effective promoter in the palladium(0)-catalyzed Suzuki cross-coupling reaction of phenylboronic acid with aryl bromides and iodide at 60-80 °C. On the basis of 13C and 31P NMR spectroscopic investigation and X-ray diffraction, it was revealed that N,N-dimethylthiocarbamate-phosphine ligand L3 might be a P,S-heterodonor bidentate ligand to palladium(0) center. 相似文献
19.
20.
Alejandro Ruiz-Rodriguez Vesna Najdanovic-Visak Zoran P. Visak Maria do Rosário Bronze Cláudia Antunes Manuel Nunes da Ponte 《Fluid Phase Equilibria》2009
This work paper presents vapour–liquid equilibrium (VLE) data for binary (CO2 + nicotine) and ternary (CO2 + nicotine + solanesol) mixtures, at 313.2 K and 6, 8 and 15 MPa. The (CO2 + nicotine) system exhibits three phases (L1L2V) in equilibrium at 8.37 MPa. It is estimated that this system most likely follows the type-III phase behaviour. In the ternary system, the presence of solanesol in the vapour phase was detected only at the pressure of 15 MPa. At this pressure, partition coefficients and separation factors for solanesol/nicotine were calculated for different initial nicotine/solanesol compositions and a strong influence of composition was found. The results were modelled using the Peng–Robinson equation of state (PR EOS) coupled with the Mathias–Klotz–Prausnitz (MKP) mixing rule (PR–MKP model). Good correlations of the binary data, particularly in the case of the (CO2 + nicotine) mixture, were obtained. However, the model could not correlate the ternary data. 相似文献