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.
First order solvolysis rates of the trans-dichlorobis (N-methylethylenediamine) cobalt(III) ion have been measured over a wide range of solvent compositions and temperatures in water–propan-2-ol and water–acetonitrile mixtures. The rate of solvolysis is faster in the former mixtures rather than the latter. Plots of log(rate constant) versus the reciprocal of the dielectric constant of the co-solvent, and also versus the Grunwald–Winstein Y-values are non-linear for both co-solvents; this non-linearity is derived from a large differential effect of solvent structure between the initial and transition states. However, extrema in the variation of enthalpy H and entropy S of activation correlate well with the extrema in physical properties of the mixtures which are related to changes in solvent structure. Linear plots of H versus S were obtained and the isokinetic temperature indicates that the reaction is entropy controlled. The application of a free-energy cycle shows that changes in solvent structure affect the pentacoordinated cobalt(III) ion in the transition state more than the hexacoordinated cobalt(III) ion in the initial state. In addition, the stabilizing influence of changes in solvent structure is greater in propan-2-ol–water mixtures than in acetonitrile–water mixtures, and the difference becomes greater as the mole fraction, x2 of the organic co-solvent increases.  相似文献   

2.
3.
4.
Solvent effects on the initial and transition states for the solvolysis of the trans-dichlorobis-(N-methylethylenediamine)cobalt(III), (meen), complex have been investigated in the 25–55 °C range in aqueous DMSO mixtures, of varying solvent composition up to 60% by vol. The log of the first order rate constant, k, varies non-linearly with the reciprocal of the dielectric constant at the same temperature, due to differential solvation of the initial and transition states. The changes in the enthalpy, H , and entropy, S , of activation with the mole fraction of the co-solvent show extrema at the composition range where the change in solvent structure occurs. The application of a free energy cycle to the process of the initial state going to the transition state suggests that the effect of solvent structure on the complex ion in the transition state dominates the initial state and that this effect increases as the mole fraction of co-solvent increases.  相似文献   

5.
The substitution reaction of the Pt(IV) complex [PtCl4(bipy)] with guanosine-5??-monophosphate (5??-GMP) was studied by UV?CVis spectrophotometry. This reaction was investigated under pseudo-first-order conditions at 37?°C in 25?mM Hepes buffer (pH?=?7.2) in the presence of 10?mM NaCl to prevent the hydrolysis of the complex. The substitution of chlorides in [{trans-Pt(NH3)2Cl}2(??-1,2-bis(4-pyridyl)ethane)](ClO4)2 (Pt3) complex by 5??-GMP was followed by 1H NMR spectroscopy under second-order conditions. Very similar values for the rate constants of both substitution steps were obtained. The Pt(IV) complexes, [PtCl4(bipy)] and [PtCl4(dach)], as well as dinuclaer Pt(II) [{trans-Pt(NH3)2Cl}2(??-pyrazine)](ClO4)2 (Pt1), [{trans-Pt(NH3)2Cl}2(??-4,4??-bipyridyl)](ClO4)2?·?DMF (Pt2) and [{trans-Pt(NH3)2Cl}2(??-1,2-bis(4-pyridyl)ethane)](ClO4)2 (Pt3) complexes, displayed potent cytotoxic activity against human ovarium carcinoma cell line TOV21G and lower activity toward human colon carcinoma HCT116 cell line at the same concentrations. Our data indicate that these platinum complexes could be explored further, as potential therapeutic agents for ovarian cancer.  相似文献   

6.
The visible spectra of VO(acac-Cl)2 and VO(acac-Br)2 have been measured in several solvents with a wide range of donor properties. Spectra in the non-coordinating solvent benzene show little difference from those of other vanadyl ß-diketonates. Results for other solvents show a linear relationship between the energy differences of the first two visible bands, D1,2 and the solvent Donor Numbers. The smaller sensitivity of D1,2 for these species to Donor Number (relative to that for VO(acac)2) has been interpreted in terms of a decreased tendency of the substituted species to undergo reaction with donor molecules.  相似文献   

7.
The complex trans-[Co(dmen)2Cl2]Cl (dmen=N,N-dimethylethylenediamine) has been prepared and characterized by elemental analysis, u.v.-vis. and i.r. spectra. The kinetics of the primary aquation of trans-[Co(dmen)2Cl2] in H2O, H2O–MeOH and H2O–i-PrOH have been examined over a wide range of solvent compositions and temperatures (40–55°C). Plots of rate constants (log k) versus the reciprocal of the dielectric constant of the medium (Ds–1) and Grunwald–Winstein values of the solvent (Y) were found to be non-linear. The variation of enthalpies (H) and entropies (S) of activation with solvent composition has been determined. Plots of H or S versus the mole fraction of each solvent exhibit extrema at x2=ca. 0.16 and 0.27 for MeOH and at x2=ca. 0.03 and 0.14 for i-PrOH. Furthermore, the cycle relating the free energy of activation in H2O to that in H2O–co-solvent shows that the stabilizing influence of the changes in the solvent structure is greater on the emergent five-coordinate cation in the transition state than that on the complex ion in the initial state, with the difference becoming greater as the mole fraction of the co-solvent increases.  相似文献   

8.
The static permittivity and polarizability of methanol–water mixtures as functions of composition at 20°C are calculated on the basis of models proposed earlier.  相似文献   

9.
《Polyhedron》1987,6(11):1993-1997
The reaction of cobalt(II) with 4-(5′-methyl-3′-isoxazolylazo)-resorcinol (MIAR) in 4% v/v ethanol-water medium at I = 0.1 M (NaClO4) was investigated spectrophotometrically. Graphical and numerical calculation methods were used to establish the equilibria in solution and to evaluate the stability constant of the complexes formed (log β101 = 7.48±0.06, log β111 = 12.77 max 12.99, log β102 = 16.41±0.07). The optimum conditions for the spectrophotometric determination of Co(II) with MIAR were established and the method applied to its determination in some low alloy steels and hydrofining catalysts.  相似文献   

10.
The dissolution enthalpies of NaI in the mixtures of methanol with 1,2-alkanediols (1,2-propanediol, 1,2-butanediol, 1,2-pentanediol) and with ??,??-alkanediols (1,3-propanediol, 1,4-butanediol, 1,5-pentanediol), as well NaI in the mixtures of water with 1,3-propanediol and 1,2-pentanediol, were determined at 298.15?K. The energetic effect of interactions between the investigated alkanediols and NaI in methanol and in water was calculated using the enthalpic pair interaction coefficients (h xy ) model. These results along with the other data concerning the NaI?Cnon-electrolyte pairs taken from our earlier reports and from the literature were analyzed with respect to the effect of the non-electrolyte properties on the variations of the h xy values. The group contributions illustrating the interactions of NaI with selected functional groups in non-electrolyte (alkanediol and alkanol) molecules, namely: CH2 and OH groups were calculated and discussed.  相似文献   

11.
The mole fraction solubility of phenacetin (PNC) in methanol + water binary solvent mixtures at 298.15 K was determined along with density of the saturated solutions. All these solubility values were correlated with the Jouyban–Acree model. Preferential solvation parameters of PNC by methanol (δx1,3) were derived from their thermodynamic solution properties using the inverse Kirkwood–Buff integrals (IKBI) method. δx1,3 values are negative in water-rich mixtures but positive in methanol mole fraction of >0.32. It is conjecturable that in the former case the hydrophobic hydration around non-polar groups of PNC plays a relevant role in the solvation. The higher solvation by methanol in mixtures of similar cosolvent compositions and methanol-rich mixtures could be explained in terms of the higher basic behaviour of methanol.  相似文献   

12.
Activity coefficients of rubidium chloride and cesium chloride in methanol–water mixed solvent systems were determined by electromotive force (EMF) measurements at 298.15 K, in the range 0–40% (wt.%) methanol. For our work, the cells:
Rb-ISE | RbCl (m), methanol (Y), water (1 − Y) | Cl-ISE;  Cs-ISE | CsCl (m), methanol (Y), water (1 − Y) | Cl-ISE
were used to obtain the emf data. The rubidium and cesium ion-selective electrodes were prepared by ourselves, and they exhibited reasonably good Nernst response. Then the experimental data were calculated by both Pitzer and Pitzer–Simonson–Clegg models. We got the corresponding parameters of two models, the standard potentials and the activity coefficients of rubidium chloride and cesium chloride. Moreover, the standard Gibbs free energies of transference of RbCl or CsCl from water to methanol–water mixtures had been discussed as well. After the comparative study, it was found that the Pitzer–Simonson–Clegg model was superior in this work.  相似文献   

13.
《Polyhedron》1986,5(11):1733-1740
Reaction of protonated 2,2′-bipyridine (bpy) with octacyanometalates(IV and V), [M(CN)8]n (M = Mo or W, n = 3 or 4), gave the following complexes: {(bpyH)3[M(CN)8]·4H2O}, {(bpyH)3(H3O)[M(CN)8]·H2O}, {(bpyH2)2[W(CN)8]·3H2O} and {(bpyH2)K2[W(CN)8]·2H2SO4·7H2O}. These salts were characterized by electron absorption and reflectance spectrometry, IR, Raman and ESR spectrometry, thermo gravimetry and differential thermal analysis, cyclic voltammetry and potentiometry. The solubility of the salts in water and some polar organic solvents has been measured. The intensively coloured salts of molybdenum(IV) and tungsten(IV) have been discussed in terms of the bpyH+-[M(CN)8]4− ion pairs, which exhibit an outer-sphere electron transfer between adjacent redox sites.  相似文献   

14.
The solvent influence on the reduction kinetics of methyl violet with iodide in binary mixture of aqueous isopropanol was investigated spectrophotometrically. The absorption spectra of methyl violet were recorded in water, aqueous isopropanol and absolute isopropanol. In these solvents λmax was in the range from 580.5 to 582.5 nm. The CNIBS/R-K model was used to calculate the solvatochromic parameters in a binary mixture; polynomial equation was also applied to describe the experimental data. The transition energies (ET) were calculated. They show bathochromic shift with the decrease in the polarity of the solvent. The temperature was varied from 298–318 K, while the pH of the reaction was maintained at 4.99 and 6.00. The reduction reaction was found to be first order by potassium iodide and zero order by methyl violet. The thermodynamic parameters were also evaluated to support the kinetic data.  相似文献   

15.
16.
17.
《Polyhedron》1987,6(2):285-288
Some new U(VI) and Ce(IV) complexes of 1-(2′-hydroxybenzyl)-2-(2′-hydroxyphenyl)-benzimidazole have been prepared and characterized by spectrg magnetic and conductance studies. IR spectral data suggests that the ligand in all the complexes is monodenate through the tertiary nitrogen and that the phenolic oxygen is free from coordination. Conductivity measurements indicate that the nitrate and acetate complexes of U(VI) are non-electrolytes, whereas the nitrate complex of Ce(IV) is a 1:1 electrolyte.  相似文献   

18.
The complexation reactions between some rare earth metal cations (Ln; Y3+, La3+ and Ce3+) with 18-crown-6 (18C6), dicyclohexyl-18-crown-6 (DC18C6), benzo-18-crown-6 (B18C6) and decyl-18-crown-6 (Dec18C6), have been studied in methanol–acetonitrile (MeOH–AN) and methanol–water (MeOH–H2O) binary mixtures using a competitive spectrophotometric method. 2-(2-thiazolylazo)-4-methyl phenol (TAC or L) was used as colorimetric complexant. It was found that the selectivity order of TAC for Ln cations is highly changed with changing the composition of the mixed solvents. Moreover, as the concentration of acetonitrile increases in MeOH–AN binary mixture, the stability of Ln–TAC complexes increases and passes through a maximum at a certain mole fraction of acetonitrile. In addition, the stability of Ln–crown ether complexes increases with increasing the concentration of methanol in MeOH–H2O and acetonitrile in MeOH–AN binary solutions. A non linear behaviour was observed for variation of stability constants of all complexes versus the composition of the mixed solvents. The results show that 18C6 generally forms more stable complexes with La3+ and Ce3+ cations than DC18C6 in methanol and MeOH–H2O binary mixtures, while this sequence is reversed in the methanol-acetonitrile binary mixtures which are rich with respect to acetonitrile.  相似文献   

19.
《Fluid Phase Equilibria》2004,219(2):257-264
A modification of the solvation model of Ohe is proposed for the calculation of vapor–liquid equilibria (VLE) in alcohol–water–salt systems. The modified method employs the Bromley equation to calculate the activity of water in salt solutions, and a one-parameter empirical expression to calculate the activity of the alcohol. The single parameter is obtained by fitting ternary alcohol–water–salt data. The method is simple to use and does not require data on the vapor-pressures of alcohol–salt mixtures that are seldom available in the literature. Experimental data for 17 salts in 36 alcohol–water–salt systems, covering a temperature range from 298 to 375 K, and salt concentrations up to about 8 m, were correlated using the new approach. In all, 69 data sets and 1045 data points were correlated satisfactorily. The method was also used to predict VLE in four ternary alcohol–alcohol–salt systems and one quaternary alcohol–alcohol–water–salt system with satisfactory results.  相似文献   

20.
Potentiometric studies of the interaction of (Me2Sn)2+ and (Me3Sn)+ with 5′-guanosine monophosphate [(5′-HGMP)2?, abbreviated as (HL-1)2?] and guanosine [(HGUO), abbreviated as (HL-2)] in aqueous solution (I = 0.1 mol·dm?3 KNO3, 298.15 ± 0.1 K) were performed, and the speciation of various complex species was evaluated as a function of pH. The species that exist at physiological pH ~7.0 are Me2Sn(HL-1)/[Me2Sn(HL-2)]2+ (87.0/88.8 %), [Me2Sn(HL-1)(OH)]?/[Me2Sn(HL-2)(OH)]+ (3.0/0 %) and [Me2Sn(HL-1H?1)]/[Me2Sn(HL-2H?1)]2+ (9.4/6.6 %) for 1:1 dimethyltin(IV):5′-guanosine monophosphate/dimethyltin(IV): guanosine systems, whereas for the corresponding 1:2 systems, the species are Me2Sn(HL-1)/[Me2Sn(HL-2)]2+ (44.0/92.0 %), [Me2Sn(HL-1H?1)]/[Me2Sn(HL-2H?1)]2+ (5.0/6.0 %), Me2Sn(OH)2 (49.0/0 %), [Me2Sn(HL-1)(OH)]?/[Me2Sn(HL-2)(OH)]+ (1.5/2.0 %), and [Me2Sn(OH)]+ (1.0/0 %). For 1:1 trimethyltin(IV):5′-guanosine monophosphate/trimethyltin(IV):guanosine systems, only [Me3Sn(HL-1)]?/[Me3Sn(HL-2)]+ (99.9 %) are found at pH = 7.0, whereas for 1:2 systems, [Me3Sn(HL-1)]?/[Me3Sn(HL-2)]+ (49.8/100 %), Me3Sn(OH) (15.0/0 %) and [Me3Sn(HL-1)(OH)]2?/Me3Sn(HL-2)(OH) (0.2/0 %) are the species found. No polymeric species were detected. Beyond pH = 8.0, significant amounts of [Me2Sn(OH)]+, Me2Sn(OH)2, [Me2Sn(OH)3]? and Me3Sn(OH) are formed. Multinuclear (1H, 13C and 119Sn) NMR studies at different pHs indicated a distorted octahedral geometry for the species Me2Sn(HL-1)/[Me2Sn(HL-2)]2+ in dimethyltin(IV)-(HL-1)2?/(HL-2) systems and a distorted trigonal bipyramidal/distorted tetrahedral geometry for the species [Me3Sn(HL-1)]?/[Me3Sn(HL-2)]+ in trimethyltin(IV)-(HL-1)2?/(HL-2) systems.  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