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 thermodynamic parameters ΔH0, ΔG0 and ΔS0 – and thereby the equilibrium constants – for the complexation of the carrier antibiotics nigericin and monensin with sodium and potassium ions in methanol at 25°C have been determined by microcalorimetry. The results are discussed in terms of the nature of the interaction between ligands and cations.  相似文献   

2.
张洪林  孔哲  闫咏梅  李干佐  于丽  李真 《化学学报》2007,65(10):906-912
N,N-二甲基乙酰胺(DMA)/长链醇非水溶液体系中, 利用微量量热仪, 研究阴离子表面活性剂十二烷基羧酸钠(SLA)、十二烷基硫酸钠(SDS)的临界胶束浓度(CMC)和热力学函数. 本文在十二烷基羧酸钠, 十二烷基硫酸钠的N,N-二甲基乙酰胺溶液中, 分别加入长链醇(庚醇、辛醇、壬醇、癸醇), 测定体系的热功率-时间曲线. 借助热力学理论, 由测得曲线, 进一步得到临界胶束浓度和热力学函数(ΔHm0, ΔGm0和ΔSm0). 讨论了温度、醇的碳原子数目、醇的浓度与热力学参数之间的关系. 结果表明, 对十二烷基羧酸钠或十二烷基硫酸钠的DMA溶液, 在含有相同浓度的各种醇的体系中, CMC, ΔHm0和ΔSm0的值随着温度的升高而增加, 而ΔGm0的值随着温度的升高而降低. 在相同温度及相同浓度的醇体系中, CMC, ΔHm0Gm0和ΔSm0的值都随着醇中碳原子数目的增加而降低. 在相同温度及相同醇的体系中, CMC, ΔGm0的值随着醇的浓度的增加而增大, 而ΔHm0, ΔSm0的值随着醇的浓度的增加而减少.  相似文献   

3.
The chromatographic behaviors of proteins on iminodiacetic acid (IDA) column with and without immobilized metal ion were examined in detail. Comparing the effects of pI, solution pH, and salt concentration on retention of proteins in cation‐exchange chromatography (CEC) and metal chelate affinity chromatography (MCAC), the retention mechanism of proteins was investigated in MCAC. By aid of observing the retention characteristics of proteins on naked IDA and metal chelate columns in high concentration salt‐out salt solution, the hydrophobic interaction in MCAC and the influence of metal ion on it were proved. In terms of the comparison of the thermodynamics of proteins in CEC and MCAC, the thermostability, the conformational change entropy Δ(ΔS0) and enthalpy Δ(ΔH0), compensation temperature β, the driving force and caloritic effect of proteins in MCAC were discussed. The identity of retention mechanism at protein thermal denaturation in CEC and MCAC was demonstrated by using the compensation relationship between ΔH0 and ΔS0.  相似文献   

4.
At different temperatures, the interactions between imidacloprid (IMI) and bovine serum albumin (BSA) were investigated with a fluorescence quenching spectrum, a synchronous fluorescence spectrum, a three-dimensional fluorescence spectrum and an ultraviolet-visible spectrum. The average values of bonding constants (KLB: 3.424 × 10^4 L,mol^-1), thermodynamic parameters (△H: 5.188 kJ,mol^-1, △G^(○—):-26.36 kJ,mol^-1, △S: 103.9 J,K^-1,mol^-1) and the numbers of bonding sites (n: 1.156) could be obtained through Stern-Volmer, Lineweaver-Burk and ther- modynamic equations. It was shown that the fluorescence of BSA could be quenched for its reactions with IMI to form a certain kind of new compound. The quenching belonged to a static fluorescence quenching, with a non-radiation energy transfer happening within a single molecule. The thermodynamic parameters agree with △H〉 0, △S〉0 and△G^(○-)〈0, suggesting that the binding power between IMI and BSA should be mainly a hydrophobic interaction.  相似文献   

5.
The aerial oxidation kinetics of hydroquinone in a freshly prepared developer solution at different temperatures and pHs has been studied. The activation parameters, Ea, ΔG# , ΔS# , ΔH# and enthalpy of formation of activated complex, ΔHfo(X# ), are determined. The large negative value of free energy of activation ΔG# proves that hydroquinone extremely tends to be oxidized by air at optimum temperature (20℃) and optimum pH (10.5) and converts to the activated complex semiquinone. It was also found that if the pH of the developer solution is increased from 9.3 to 10.5 the reaction rate will increase by a factor of 2.  相似文献   

6.
The critical micelle concentration (CMC) of two kinds of anionic surfactant (including sodium laurate (SLA) and sodium dodecyl sulfate (SDS)) in mixed alcohol and N, N‐dimethyl formamide solvent (DMF) were investigated through measuring power‐time curves by titration microcalorimetry. From data of the lowest point and the area of the power‐time curves, their CMC and ΔH m 0 can be obtained. According to standard thermodynamic equations, ΔG m 0 and ΔS m 0 also can be calculated. For different surfactant, the influences of the carbon number and the concentration of alcohol on the CMC and standard thermodynamic functions are different in DMF polar medium. These thermodynamic functions for micelle formation can be further interpreted.  相似文献   

7.
Abstract— This study was undertaken to further investigate the way in which the counter anion controls the Λmax of the absorption spectrum of compounds similar to N-retinylidene-n-butylammonium salts (NRBA). The following relationship had been found: ΔE =ΔEo -F d0e2/εd2; here ΔE is the observed excitation energy, e the charge on the electron, ε the dielectric constant, d0 a constant and d the distance between centers of opposite charge as estimated from crystallographic radii. Resonance theory implies that ΔEo should be of the same numerical value as the corresponding carbonium ion which can be generated readily from the corresponding alcohol. The C22SB analog of NRBA was prepared and then converted to the halide salts. The Δmax of these salts was determined in several halohydrocar-bon solvents, and ΔEo was determined by least squares for each solvent. The average value of ΔEo was found to be 653 nm, while the Λmax, for the carbonium ion was previously found to be 644 nm. The results are supportive of previous work.  相似文献   

8.
Interaction between adsorbed hydrogen and the coordinatively unsaturated Mg2+ and Co2+ cationic centres in Mg‐MOF‐74 and Co‐MOF‐74, respectively, was studied by means of variable‐temperature infrared (VTIR) spectroscopy. Perturbation of the H2 molecule by the cationic adsorbing centre renders the H? H stretching mode IR‐active at 4088 and 4043 cm?1 for Mg‐MOF‐74 and Co‐MOF‐74, respectively. Simultaneous measurement of integrated IR absorbance and hydrogen equilibrium pressure for spectra taken over the temperature range of 79–95 K allowed standard adsorption enthalpy and entropy to be determined. Mg‐MOF‐74 showed ΔH0=?9.4 kJ mol?1 and ΔS0=?120 J mol?1 K?1, whereas for Co‐MOF‐74 the corresponding values of ΔH0=?11.2 kJ mol?1 and ΔS0=?130 J mol?1 K?1 were obtained. The observed positive correlation between standard adsorption enthalpy and entropy is discussed in the broader context of corresponding data for hydrogen adsorption on cation‐exchanged zeolites, with a focus on the resulting implications for hydrogen storage and delivering.  相似文献   

9.
金属Pt表面水蒸汽分子吸附的量子力学计算   总被引:1,自引:0,他引:1  
基于电子与振动近似方法和密度泛函B3LYP理论, 氧和氢原子选择6-311G**基函数, Pt选择赝势基组LanL2DZ, 优化得到Pt-OH2结构和微观性质, 稳态结构Pt-H2O分子中, Pt与H2O不在同一平面, Pt倾向于与O原子结合. 计算了100~898.15 K温度下, 水蒸汽分子在Pt表面吸附反应的热力学函数值和平衡压力, 拟合得到ΔS0H0, ΔG0, ln p与温度的函数关系. 室温以上ΔG0>0 kJ•mol-1, 水蒸汽分子在Pt表面不能稳定吸附; 200 K以下, ΔG0<0 kJ•mol-1, 能够稳定吸附. 计算了不同温度下水蒸汽分子在Pt表面发生解离反应的ΔG0和平衡压力, 室温以上ΔG0>0 kJ•mol-1. 100~898.15 K温度下, 水蒸汽分子在Pt表面不容易发生解离, 实际反应过程中以完整分子形式参与反应.  相似文献   

10.
Is There a Minimum Electrophilicity Principle in Chemical Reactions?   总被引:1,自引:0,他引:1  
NOORIZADEH  Siamak 《中国化学》2007,25(10):1439-1444
For 25 simple reactions, the changes of the hardness (△η), polarizability (△α) and electrophilicity index (△ω) and their cube-roots (△η^1/3, △α^1/3, △ω^1/3) were calculated. It is shown that although the Maximum Hardness and Minimum Polarizability Principles are not valid for all reactions, but in most cases △ω^1/3〈0, whereas we always find △ω〈0. Our observation implies to this fact that for those chemical reactions in which the number of moles decreases or at least remains constant, the most stable species (reactants or products) have the lowest sum of electrophilicities. In other words "the natural direction of a chemical reaction is toward a state of minimum electrophilicity". This fact may be called Minimum Electrophilicity Principle (MEP).  相似文献   

11.
Densities have been measured for the electrolyte (NaCl, NaBr and NaI)‐monosaccharide (D ‐mannose and D‐ribose)‐water solutions at 298.15 K. These data have been used to calculate the apparent molar volumes of the saccharides (VΦ,S) and electrolytes (VΦ,E) in the studied solutions. Infinite dilution apparent molar volumes, VΦ,S0 and VΦ,E0, have been evaluated, together with the standard transfer volumes of the saccharides (ΔtVS0) from water to aqueous electrolyte solutions and those of the electrolytes (ΔtVE0) from water to aqueous saccharide solutions. It was shown that both the ΔtVS0 and ΔtVE0 values are positive and increase with increasing molalities of sodium halides and saccharides, respectively. Overall, the ΔtVS0 and ΔtVE0 values have the order of NaCl > NaBr > NaI except for NaI‐ribose and NaI‐ribose. Volumetric interaction parameters for the electrolyte‐monosaccharide pairs in water were obtained and interpreted by the stereochemistry of the monosaccharide molecules and the structural interaction model.  相似文献   

12.
The density and the viscosity data have been used to determine the thermodynamic activation parameters, free energies (ΔG ?), enthalpies (ΔH ?) and entropies (ΔS ?), for viscous flow of the systems; water (W) + ethylenediamine (ED), W + trimethylenediamine (TMD) and W + N,N-dimethyltrimethylenediamine (DMTMD) in the temperature range of 303.15–323.15 K over the composition range of 0 ≤ X 2 ≤ 0.45, where X 2 is the mole fraction of diamines. On addition of diamines to water, ΔG ?, ΔH ? and ΔS ? values increase sharply, pass through a maximum and then decline. The heights of maximum in the ΔG ? versus X 2 curve vary as, W + DMTMD > W + TMD > W + ED. For all systems, the excess properties, ΔG ? E , ΔH ? E and ΔS ? E are positive. The observed increase in thermodynamic values may be due to combined effect of hydrophobic hydration of diamines and water–diamine interaction as a result of hydrophilic effect.  相似文献   

13.
In this study activated carbon was used for the removal of thiram from aqueous solutions. Adsorption experiments were carried out as a function of time, initial thiram concentration and temperature. Equilibrium data fitted well to the Freundlich and Langmuir equilibrium models in the studied concentration range. Adsorption kinetics followed a pseudo second‐order kinetic model rather than pseudo first‐order model. The results from kinetic experiments were used to describe the adsorption mechanism. Both boundary layer and intraparticle diffusion played important role in the adsorption mechanism of thiram. Thermodynamic parameters (ΔG0, ΔH0, and ΔS0) were determined and the adsorption process was found to be an endothermic one. The negative values of ΔG0 at different temperatures were indicative of the spontaneity of the adsorption process.  相似文献   

14.
The equilibrium constant for the reaction C6H5NHCOCl = C6H5NCO + HCl in chlorobenzene solution is K = 0.14 mole/kg at 70°. The approximate values of the enthalpy and the entropy of the reaction are ΔH0 = 12 kcal and ΔS0 = 31 cal/deg.  相似文献   

15.
《Analytical letters》2012,45(16):2693-2710
Abstract

The binding interaction of the terazosin hydrochloride and prazosin hydrochloride with bovine serum albumin (BSA) was studied by spectrofluorimetry. Both of these two compounds quenched the fluorescence of BSA. The thermodynamic parameters (ΔH 0, ΔS 0 and ΔG 0) obtained from the fluorescence data measured at two different temperatures showed that the binding of terazosin hydrochloride to BSA involved hydrogen bonds and that of prazosin hydrochloride to BSA involved hydrophobic and electrostatic interactions. In this work, the competitive interaction of the terazosin hydrochloride and prazosin hydrochloride with BSA was studied by three-way excitation-emission fluorescence with the aid of parallel factor analysis (PARAFAC).  相似文献   

16.
Carboxylate and salicylic OH coordinate bonding as well as intramolecular and intermolecular hydrogen bonding of bis-3,5-diisopropylsalicylatozinc(II), [ZnII(3,5-DIPS)2], with Lewis bases were studied to determine mechanisms accounting for antioxidant reactivity of ZnII(3,5-DIPS)2. Apparent thermodynamic parameters: K eq, ΔS 0, ΔH 0, and ΔG 0 were determined for these equilibria with bonding of two molecules of dimethyl sulfoxide-d6 (DMSO) or ethyl acetate-d8 (EA) to the ZnII using NMR and FTIR. We conclude that addition of two equivalents of DMSO or EA to non-polar solutions of ZnII(3,5-DIPS)2 results in bonding of DMSO or EA to ZnII via sulfoxide or ester carbonyl oxygen atoms with ternary complex formation, leading to weakening of carboxylate and salicylic OH coordinate bonding to ZnII and strengthening intramolecular hydrogen bonding between protons of salicylic OH groups and carboxylate oxygens. Subsequent addition of two or three additional equivalents of DMSO or EA leads to intermolecular hydrogen bonding between protons of salicylic OH groups.  相似文献   

17.
The reaction between chromium(VI) and L-ascorbic acid has been studied by spectrophotometry in the presence of aqueous citrate buffers in the pH range 5.69–7.21. The reaction is slowed down by an increase of the ionic strength. At constant ionic strength, manganese(II) ion does not exert any appreciable inhibition effect on the reaction rate. The rate law found is where Kp is the equilibrium constant for protonation of chromate ion and kr is the rate constant for the redox reaction between the active forms of the oxidant (hydrogenchromate ion) and the reductant (L-hydrogenascorbate ion). The activation parameters associated with rate constant kr are Ea = 20.4 ± 0.9 kJ mol?1, ΔH = 17.9 ± 0.9 kJ mol?1, and ΔS=?152 ± 3 J K?1 mol?1. The reaction thermodynamic magnitudes associated with equilibrium constant Kp are ΔH0 = 16.5 ± 1.1 kJ mol?1 and ΔS0 = 167 ± 4 J K?1 mol?1. A mechanism in accordance with the experimental data is proposed for the reaction. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
The thermodynamic activation parameters, enthalpies, ΔH?, free energies, ΔG , and entropies, ΔS?, for viscous flow of the systems, water (W)?+?n-butylamine (NBA), W?+?sec-butylamine (SBA) and W?+?tert-butylamine (TBA), have been determined by using the density and the viscosity data. These properties and their excess values have been represented graphically against their composition. With respect to the composition, ΔG show a typical behaviour for all the systems – a fast rise in the water-rich region with a maximum followed by the values that decline up to the pure state of amines. The ΔH? and ΔS? versus composition curves follow the similar trend. For all systems the excess properties, ΔG ≠E, ΔH ?≠E and ΔS?≠E are characterized by sharp maxima in the water-rich region, which are thought to be mainly due to the hydrophobic hydration and the hydrophilic effect.  相似文献   

19.
The dependence of indium trichloride saturated and unsaturated vapor pressure on temperature was studied in the range of 630–950 K by static methods using a quartz membrane zero‐manometer and taking into account the volume of its working chamber and substance mass. The thermodynamic data on the process of dissociation of dimeric molecules and sublimation of monomer and dimer from solid indium trichloride were calculated: ΔH0subl InCl3(g)298 = 155.3 ± 6.2 kJ · mol–1; ΔS0subl InCl3(g)298 = 199.5 ± 7.9 J · mol–1 · K–1; ΔH0subl In2Cl6(g)298 = 159.3 ± 6.2 kJ · mol–1; ΔS0subl In2Cl6(g)298 = 207.1±3.8 J · mol–1 · K–1; ΔH0dis In2Cl6(g)298 = 152.6 ± 5.5 kJ · mol–1 and ΔS0dis In2Cl6(g)298 = 171.6 ± 5.2 J · mol–1 · K–1. The saturated vapor over solid indium trichloride consists mainly of a mixture of monomeric and dimeric molecules (InCl3 and In2Cl6), and the content of the latter is slightly growing with increasing temperature.  相似文献   

20.
Interpolymer complexes of methacrylic acid-acrylamide copolymer have been prepared by the interaction of its comonomer units with two different homopolymers, and also with a single homopolymer. The stability constants and related thermodynamic parameters (e.g., ΔH0 and ΔS0) of the mixed complex were found to be significantly different from the single homopolymer complexes. This has been attributed to neighboring group influence and destabilization of various interacting forces at different temperatures. © 1994 John Wiley & Sons, Inc.  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