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.
Partial volumes $\bar V^0$ of amino acids in aqueous NH4Cl and NaCl solutions are discussed. The salts have different effects on water structure. The contributions of the charged NH 3 + and COO? groups of amino acids are found. Structural characteristics of hydrated complexes are calculated: partial volumes of water inside and outside the hydration sphere and hydration numbers. The same value of $\bar V^0$ (NH 3 + , COO?) is achieved at a higher NH4Cl concentration. The two salt systems with the same $\bar V^0$ (NH 3 + , COO?) have similar values of the partial volumes of water and hydration numbers.  相似文献   

2.
The apparent volumes of the salts in the systems H2O-NH4Cl (298 K) and H2O-NH4NO3 (273 K, 298 K, and 323 K) are reproduced with an accuracy of 0.03–0.01 cm3/mol by the equation ? = ?0 + Aw 2 0.5 + Bw 2, where w 2 is the salt content (mass fractions). The study shows that there is a correspondence between the critical (for determining the hydration number) structural parameters-the intrinsic volume of the electrolyte and the volume of water in ion hydration shells-and the limiting (at w 2 = 1) partial molar volumes of the components. The hydration numbers at infinite dilution are 6.9 for NH4Cl at 298 K and 9.1, 6.7, and 6.4 for NH4NO3 at 273 K, 298 K, and 323 K. The water volume in ion hydration shells decreases in the sequence: No 3 ? , Cl?, and NH 4 + . The hydration numbers decrease with increasing salt concentration. The study shows that within a simpler model ? = ?0 + aw 2 0.5 , the hydration numbers are temperature independent.  相似文献   

3.
许莉  丁成荣  林环  王旭  林瑞森 《化学学报》2007,65(23):2797-2801
应用精密数字密度计测定了298.15 K时甘氨酸、L-丙氨酸或L-缬氨酸与不同组成的木糖醇-水混合溶剂构成的三元系溶液的密度, 计算了氨基酸的表观摩尔体积、极限偏摩尔体积、迁移偏摩尔体积和水化数. 根据结构水合作用模型讨论了迁移偏摩尔体积和水化数的变化规律. 结果表明, 氨基酸两性离子部分与木糖醇羟基间的相互作用占主导地位, 且随木糖醇浓度的增大而增大. 氨基酸在木糖醇水溶液中的迁移偏摩尔体积为正值, 甘氨酸的迁移偏摩尔体积大于L-丙氨酸和L-缬氨酸的. 氨基酸在木糖醇水溶液中的水化数随溶液中木糖醇浓度的增加而减小.  相似文献   

4.
The dependence of the standard partial volumes of glycine, α-alanine, and serine on the ionic strength of aqueous sodium chloride and sulfate solutions is modeled by the extended Masson equation: {ie534-1}. The error of less than 0.2 cm3/mol is a result of using five values of the A and B parameters: the two values of A are determined by the type of salt and the three values of B by the type of amino acid. A new variation of the additive-group approach is proposed for {ie534-2}. The partial volumes of the CH3 group (α-alanine) and the CH2 group (serine) are found not to depend on the salt concentration. The partial volume of the CH2 group of glycine grows with concentration. The structural characteristics of the hydrated complexes of the NH 3 + and COO? groups are calculated: the hydration numbers, the molar volumes of water inside and outside the hydration sphere, and the intrinsic volume of NH 3 + in COO? in solution. Given the same ionic strength, the aqueous sodium sulfate solution produces a somewhat stronger dehydration of the charged groups.  相似文献   

5.
Ion hydration in aqueous solutions of KOH, KCl, KI, and KIO3 was studied by refractometry. The sum of the hydration numbers of the potassium ion and each anion in these compounds and the hydration numbers of OH?, Cl?, I?, and IO 3 ? anions were determined by the Lorentz-Lorenz equation at different salt concentrations. The model suggests that the polarizability of the hydrated ion is proportional to the cube of the ion radius and the volume of the hydrated ion equals the sum of the nonhydated ion and hydration shell volumes. The hydration numbers of the anions calculated by the proposed procedure increase with their radii in the sequence: OH?, Cl?, I?, and IO 3 ? .  相似文献   

6.
The apparent molar volumes, V φ , of L-aspartic acid, L-glutamic acid, L-lysine monohydrate and L-arginine in water and in aqueous (0.1, 0.25, 0.5 and 1.0) mol?kg?1 sodium acetate and sodium propionate, and (0.1, 0.25 and 0.5) mol?kg?1 sodium butyrate solutions have been determined at 288.15, 298.15, 308.15 and 318.15 K from density measurements. The partial molar volumes at infinite dilution, V 2 o , obtained from V φ data, have been used to calculate hydration numbers and partial molar expansibilities of amino acids in water and in the presence of the studied cosolutes at different temperatures. These parameters have been discussed in terms of various interactions between the acidic/basic amino acids and organic salts in these solutions. The effect of the hydrophobic chain length of the carboxylate ions has also been discussed.  相似文献   

7.
A new modification of the adiabatic compressibility method of investigating solvation in solutions is presented and applied to the analysis of the following structurally-related characteristics of hydrated complexes of seawater electrolytes (NaCl, KCl, MgCl2, CaCl2, Na2SO4, MgSO4) at different concentrations (0.1 to 5.0 mol⋅kg−1) and temperatures (278.15 to 308.15 K): solvation numbers (h) and their dependences on concentration, volumes of stoichiometric mixtures of ions without their hydration shells (V 2h ), compressibilities (β 1h ) and molar volumes of water in their solvation shells (V 1h ), their dependences on concentration and temperature, etc.  相似文献   

8.
With an increase in the concentration of additives, the hydration numbers of compounds decrease. Thus, in a saturated 54.6% solution, urea loses approximately 3/4 of the initial amount of water, forming an aquacomplex of the composition (NH2)2CO?H2O. In a supersaturated 44% solution, the sodium chloride aquacomplex is dehydrated by 2/3, and in a supersaturated 67% solution, sodium sulfate is dehydrated by 5/6. The density of these solutions is 1.354÷1.360 g/cm3 (44% NaCl) and 1.800÷1.849 g/cm3 (67% Na2SO4). In a saturated urea solution, NaNO3, NaCl, and Na2SO4 complexes lose 53÷55% of hydration water. It is shown that the interactions in the binary water–urea system somewhat increase the hydration number of the salts (structural hydration). The hydration water density, a structurally important characteristic, increases in the series of solutions of urea, NaNO3, NaCl, and Na2SO4. In the same series of additives, the excess volume of binary water–urea and water–salt systems becomes more negative.  相似文献   

9.
The hydration number of α-alanine in aqueous urea solutions is greater than in aqueous NaCl solutions; the ratio of the hydration numbers increases from 0.2 (m = 1) to ≈2 (m = 6). Given the same partial volumes of water, the hydration numbers of α-alanine in the two systems are close to each other.  相似文献   

10.
The ab initio quantum mechanical charge field molecular dynamics (QMCF MD) formalism was applied to simulate the bicarbonate ion, HCO3?, in aqueous solution. The difference in coordination numbers obtained by summation over atoms (6.6) and for the solvent‐accessible surface (5.4) indicates the sharing of some water molecules between the individual atomic hydration shells. It also proved the importance to consider the hydration of the chemically different atoms individually for the evaluation of structural and dynamical properties of the ion. The orientation of water molecules in the hydration shell was visualized by the θ–tilt surface plot. The mean residence time in the surroundings of the HCO3? ion classify it generally as a structure‐breaking ion, but the analysis of the individual ion‐water hydrogen bonds revealed a more complex behavior of the different coordination sites. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

11.
A great deal of information exists concerning the hydration of ions in bulk water. Much less noticeable, but equally ubiquitous is the hydration of ions holding on to several water molecules in nanoscopic pores or in natural air at low relative humidity. Such hydration of ions with a high ratio of ions to water molecules (up to 1:1) are essential in determining the energetics of many physical and chemical systems. Herein, we present a quantitative analysis of the energetics of ion hydration in nanopores based on molecular modeling of a series of basic salts associated with different numbers of water molecules. The results show that the degree of hydrolysis of basic salts in the presence of a few water molecules is significantly different from that in bulk water. The reduced availability of water molecules promotes the hydrolysis of divalent and trivalent basic ions (S2?, CO32?, SO32?, HPO42?, SO42?, PO43?), which produces lower valent ions (HS?, HCO3?, HSO3?, H2PO4?, HSO4?, HPO42?) and OH?ions. However, reducing the availability of water inhibits the hydrolysis of monovalent basic ions (CN?, HS?). This finding sheds some light on a vast number of chemical processes in the atmosphere and on solid porous surfaces. The discovery has wide potential applications including designing efficient absorbents for acidic gases.  相似文献   

12.
The relations for the calculation of the partial molar volumes of NaCl, NaBPh4, and Ph4AsCl in an aqueous urea solution are obtained. The salt characteristics are divided into ionic components. Different methods of the division are discussed. It is shown that the hydration numbers of Na+ and Cl ions decrease with increasing urea concentration; therewith, the dehydration of Cl ion occurs relatively faster.  相似文献   

13.
The electrode reaction of the Ni(II)/Ni(Hg) system occurring in 4–13.7 mol kg?1 solutions of Ca(ClO4)2 in the temperature range 20–185°C was studied by means of pulse polarography and cyclic voltammetry. From analysis of the formal potentials with respect to the ferricinium ionferrocene electrode as a function of the logarithm of the water activity, the hydration numbers of Ni(II) were found. They decrease as both the molality of Ca(ClO4)2 and the temperature increase. Limiting values of hydration numbers at both high temperature and concentration of Ca(ClO4)2 were near to 6. The electrode reaction proceeds in a step-wise manner with the Ni(I)/Ni(Hg) system controlling the rate of the overall process.The rate constants at formal potentials under different conditions are presented.The electrode reaction orders are determined with respect to the water point to the participation of Ni(H2O)4+ in the rate-determining step.  相似文献   

14.
The apparent molar volumes Vφ of glycine, alanine, valine, leucine, and lysine have been determined in aqueous solutions of 0.05, 0.5, 1.0 mol · kg−1 sodium dodecyl sulfate (SDS) and 1.0 mol · kg−1 cetyltrimethylammonium bromide (CTAB) by density measurements at T=298.15 K. The apparent molar volumes have also been determined for diglycine and triglycine in 1 mol · kg−1 SDS and CTAB solutions. These data have been used to calculate the infinite dilution apparent molar volumes V20 for the amino acids and peptides in aqueous SDS and CTAB and the standard partial molar volumes of transfer (ΔtrV2,m0) of the amino acids and peptides to these aqueous surfactant solutions. The linear correlation of V20 for a homologous series of amino acids has been utilized to calculate the contribution of the charged end groups (NH3+, COO), CH2 group and other alkyl chains of the amino acids to V20. The results on the partial molar volumes of transfer from water to aqueous SDS and CTAB have been interpreted in terms of ion–ion, ion–polar and hydrophobic–hydrophobic group interactions. The volume of transfer data suggests that ion–ion or ion–hydrophilic group interactions of the amino acids and peptides are stronger with SDS compared to those with CTAB. Comparison of the hydration numbers of amino acids calculated in the present studies with those in other solvents from literature shows that these numbers are almost the same at 1 mol · kg−1 level of the cosolvent/cosolute. Increasing molality of the cosolvent/cosolute beyond 1 mol · kg−1 lowers the hydration number of the amino acids due to increased interactions with the solvent and reduced electrostriction.  相似文献   

15.
Experimental data on the speed of propagation of ultrasound waves, density, and isobaric heat capacity in aqueous solutions of urea and urotropin have been considered. The findings have been used for calculating the molar isentropic compressibilities of solutions of the investigated substances over the temperature range 278.15 to 308.15 K. Invoking a theoretical solvation model based on the isentropic compressibility, which takes into account compressibilities of the hydrated complexes, their structural characteristics have been determined in aqueous solutions of nonelectrolytes: hydration numbers h, molar isentropic compressibility of hydrated complexes ?? h V h , molar volumes of water in a hydration shell V 1h , molar volumes of the solute without its hydration environment V 2h , and many other properties. The possibility of hydrophobic solvation has been shown for urotropin solutions and hydrophilic solvation for urea solutions.  相似文献   

16.
A coordination study of Lu(III) has been carried out for the nitrate and perchlorate salts in aqueous mixtures of acetone-d6 and Freon-12 by1H,15N and35Cl NMR spectroscopy. At temperatures lower than –90°C, proton and ligand exchange are slow enough to permit the direct observation of1H resonance signals for coordinated and free water molecules, leading to an accurate measure of the Lu(III) hydration number. In perchlorate solution, in the absence of inner-shell ion-pairing, Lu(III) exhibits a maximum coordination number of six over the allowable concentration range of study, contrasting markedly with the report of values of six to nine or greater as determined by a similar NMR method. The absence of contact ion-pairing was confirmed by35Cl NMR chemical shift and linewidth measurements. Extensive ion-pairing was observed in the nitrate solutions as reflected by the lower Lu(III) hydration numbers of two to three in these systems, the observation of two coordinated water signals, and15N NMR signals for two complexes. The1H and15N NMR spectra and the hydration number could be accounted for by the presence of (H2O)4Lu(NO3)2+ and (H2O)2Lu(NO3) 2 1+ .  相似文献   

17.
18.
The 17O-NMR spin-lattice relaxation times (T 1) of water molecules in aqueous solutions of n-alkylsulfonate (C1 to C6) and arylsulfonic anions were determined as a function of concentration at 298 K. Values of the dynamic hydration number, (S-) = nh - (tc- /tc0 - 1)(\mathrm{S}^{-}) = n_{\mathrm{h}}^{ -} (\tau_{\mathrm{c}}^{-} /\tau_{\mathrm{c}}^{0} - 1), were determined from the concentration dependence of T 1. The ratios (tc -/tc0\tau_{\mathrm{c}}^{ -}/\tau_{\mathrm{c}}^{0}) of the rotational correlation times (tc -\tau_{\mathrm{c}}^{ -} ) of the water molecules around each sulfonate anion in the aqueous solutions to the rotational correlation time of pure water (tc0\tau_{\mathrm{c}}^{0}) were obtained from the n DHN(S) and the hydration number (nh -n_{\mathrm{h}}^{ -} ) results, which was calculated from the water accessible surface area (ASA) of the solute molecule. The tc -/tc0\tau_{\mathrm{c}}^{ -}/\tau_{\mathrm{c}}^{0} values for alkylsulfonate anions increase with increasing ASA in the homologous-series range of C1 to C4, but then become approximately constant. This result shows that the water structures of hydrophobic hydration near large size alkyl groups are less ordered. The rotational motions of water molecules around an aromatic group are faster than those around an n-alkyl group with the same ASA. That is, the number of water–water hydrogen bonds in the hydration water of aromatic groups is smaller in comparison with the hydration water of an n-alkyl group having the same ASA. Hydrophobic hydration is strongly disturbed by a sulfonate group, which acts as a water structure breaker. The disturbance effect decreases in the following order: $\mbox{--} \mathrm{SO}_{3}^{-} > \mbox{--} \mathrm{NH}_{3}^{ +} > \mathrm{OH}> \mathrm{NH}_{2}$\mbox{--} \mathrm{SO}_{3}^{-} > \mbox{--} \mathrm{NH}_{3}^{ +} > \mathrm{OH}> \mathrm{NH}_{2}. The partial molar volumes and viscosity B V coefficients for alkylsulfonate anions are linearly dependent on their n DHN(S) values.  相似文献   

19.
The hydration of doubly protonated gas-phase ions of gramicidin S formed by electrospray ionization was investigated. Under “gentle” electrospray conditions, a near Gaussian distribution of (M + 2H + nH2O)2+ ions with n up to 50 can be readily formed. These extensively hydrated gas-phase ions should have structures similar to those in solution. For intermediate extents of hydration, the “naked” or unsolvated ion is present in unusually high abundance. This is attributed to a competition between solvation of the charges by water vs intramolecular self-solvation via hydrogen bonding. In addition, “magic” numbers of attached water molecules are observed for n = 8, 11, and 14. These magic numbers are attributed to favorable arrangements of water molecules surrounding the charge and surface of the peptide in the gas phase. These results are indicative of a gentle stepwise transformation from the solution-phase structure of the ion to the preferred gas-phase structure as solvent evaporates from the hydrated ions.  相似文献   

20.
测定了278.15~318.15 K(间隔10 K)下葡萄糖+HCl+水三元体系的密度, 计算了葡萄糖在盐酸(浓度0.2~2.1 mol•kg–1)中的表观摩尔体积VΦ,G、标准偏摩尔体积V0Φ,G、葡萄糖-HCl在水中的体积对相互作用参数VEG和标准偏摩尔膨胀系数(∂V0Φ,G/∂T)p. 结果表明: (1)葡萄糖在盐酸中的表观摩尔体积随葡萄糖和HCl的浓度的增加而线性增大; (2) V0Φ,G随HCl的质量摩尔浓度的增加而线性增大; (3)葡萄糖与HCl在水溶液中的体积相互作用参数VEG>0, 但数值对温度变化不甚敏感; (4)葡萄糖在水和盐酸中的V0Φ,G值随实验温度的变化关系均可表示为: V0Φ,Ga0a1(T-273.15 K) 2/3; (5) (∂V0Φ,G/∂T)p为正值且随温度的升高而减小; 在一定温度下, 其值随HCl浓度的增加而稍稍减小. 糖的水化程度随温度的升高和HCl的浓度的增加而减小. 用结构相互作用模型对葡萄糖与HCl之间的体积相互作用进行了解释.  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