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.
Using the analytical gravimetric method the solubility of glycine, dl-alanine, l-isoleucine, l-threonine, and l-serine in aqueous systems of (NH4)2SO4, at (298.15 and 323.15) K, were measured for salt concentrations ranging up to 2.0 molal.In the electrolyte molality range studied the experimental observations showed that ammonium sulfate is a salting-in agent for most of the amino acids studied. Furthermore, the change of the relative solubility with electrolyte concentration shows a maximum, which makes the representation of the data by a simple empirical correlation such as the Setschenow equation difficult. For the development and evaluation of a robust thermodynamic framework that makes it possible to more profoundly understand aqueous amino acid solutions with ammonium sulfate additional experimental information is needed.  相似文献   

2.
We report the results of a selected ion flow tube (SIFT) study of the reactions of H3O+, NO+ and O+2 with some nine carboxylic acids and eight esters. We assume that all the exothermic proton transfer reactions of H3O+ with all the acid and esters molecules occur at the collisional rate, i.e. the rate coefficients, k, are equal to kc; then it is seen that k values for most of the NO+ and O+2 reactions also are equal to or close to kc. The major ionic products of the H3O+ reactions with both the acids and esters are the protonated parent molecules, MH+, but minor channels are also evident, these being the result of H2O elimination from the excited (MH+)1 in some of the acid reactions and an alcohol molecule elimination (CH3OH or C2H5OH) in some of the ester reactions. The NO+ reactions with the acids and esters result in both ion-molecule association producing NO+M in parallel with hydroxide ion (OH) transfer with some of the acids, and parallel methoxide ion (CH3O) and ethoxide ion (C2H5O) transfer as appropriate with some of the esters. The O+2 reactions proceed by dissociative charge transfer with the production of two or more ionic fragments of the parent molecules, the different isomeric forms of both the acid and the ester molecules resulting in different product ions.  相似文献   

3.
New salt forms of the antioxidant drug emoxypine (EMX, 2‐ethyl‐6‐methylpyridin‐3‐ol) with pharmaceutically acceptable maleic (Mlt), malonic (Mln) and adipic (Adp) acids were obtained {emoxypinium maleate, C8H12NO+·C4H3O4, [EMX+Mlt], emoxypinium malonate, C8H12NO+·C3H3O4, [EMX+Mln], and emoxypinium adipate, C8H12NO+·C6H9O4, [EMX+Adp]} and their crystal structures determined. The molecular packing in the three EMX salts was studied by means of solid‐state density functional theory (DFT), followed by QTAIMC (quantum theory of atoms in molecules and crystals) analysis. It was found that the major contribution to the packing energy comes from pyridine–carboxylate and hydroxy–carboxylate heterosynthons forming infinite one‐dimensional ribbons, with [EMX+Adp] additionally stabilized by hydrogen‐bonded C(9) chains of Adp ions. The melting processes of the [EMX+Mlt] (1:1), [EMX+Mln] (1:1) and [EMX+Adp] (1:1) salts were studied and the fusion enthalpy was found to increase with the increase of the calculated lattice energy. The dissolution process of the EMX salts in buffer (pH 7.4) was also studied. It was found that the formation of binary crystals of EMX with dicarboxylic acids increases the EMX solubility by more than 30 times compared to its pure form.  相似文献   

4.
The effect of benzene carboxylic acids on the adsorption of Cd(II) (5×10−5 M) by goethite and kaolinite has been studied in 0.005 M NaNO3 at 25°C. The concentrations of phthalic (benzene-1,2-dicarboxylic acid), hemimellitic (1,2,3), trimellitic (1,2,4), trimesic (1,3,5), pyromellitic (1,2,4,5) and mellitic (1,2,3,4,5,6) acids varied from 2.5×10−5 to 1×10−3 M. Mellitic acid complexes Cd(II) strongly above about pH 3, but the other acids only at higher pH, phthalic acid forming the weakest complexes. Phthalic, trimesic and mellitic acids adsorbed strongly to goethite at pH 3, but adsorption decreased at higher pH; however, mellitic acid was still about 50% adsorbed at pH 9, by which the other two were almost entirely in solution. At 10−3 M all the acids enhanced the adsorption of Cd(II) to goethite, the higher members of the series being the most effective. The higher members of the series suppressed Cd(II) adsorption onto kaolinite, but phthalic and trimesic acids caused slight enhancement. The effects of mellitic acid on Cd(II) adsorption depended strongly on its concentration. The maximum enhancement of Cd(II) adsorption onto goethite was at 10−4 M. The greatest suppression of Cd(II) adsorption onto kaolinite was at 10−3 M, and at 2.5×10−5 M mellitic acid enhanced Cd(II) adsorption onto kaolinite at intermediate pH. The results are interpreted in terms of complexation between metal and ligand (acid), metal and substrate, ligand and substrate, and the formation of ternary surface complexes in which the ligand acts as a bridge between the metal and the surface.  相似文献   

5.
The separations of amino acids by Donnan dialysis using an ion-exchange membrane were studied. Donnan dialytic experiments were carried out using an anion-exchange membrane, glutamic acid–phenylalanine or glutamic acid–alanine mixed solutions as the feeds, and sodium hydroxide solutions as the stripping ones. The initial concentrations of the two kinds of amino acids in the feed solutions were equal and in the range of 0.5–50 mol m−3. The amino acid fluxes were measured for each feed solution. Above the feed concentration of 10 mol m−3, the glutamic acid flux was over 100 times greater than that of the other amino acid, and it was found that the Donnan dialysis was applicable to the separation of the amino acids. On the other hand, below 10 mol m−3, the amino acid fluxes varied in a complicated manner with the concentration, and below 1 mol m−3 there was little difference between the fluxes of the two amino acids.Furthermore, after soaking the membrane in solutions having the same concentrations as the feed in the Donnan dialysis, uptake of the amino acids into the membrane was also measured. By comparing the experimental results of both the flux and uptake of the amino acids, the reason why the flux varied in a complicated manner with the concentration was discussed.  相似文献   

6.
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   

7.
The effect of glycine, dl-alanine and dl-2-aminobutyric acid on the temperature of maximum density of water was determined from density measurements using a magnetic float densimeter.Densities of aqueous solutions were measured within the temperature range from T = (275.65 to 278.65) K at intervals of T = 0.50 K over the concentration range between (0.0300 and 0.1000) mol · kg−1. A linear relationship between density and concentration was obtained for all the systems in the temperature range considered.The temperature of maximum density was determined from the experimental results. The effect of the three amino acids is to decrease the temperature of maximum density of water and the decrease is proportional to molality according to Despretz equation. The effect of the amino acids on the temperature of maximum density decreases as the number of methylene groups of the alkyl chain becomes larger. The results are discussed in terms of (solute + water) interactions and the effect of amino acids on water structure.  相似文献   

8.
Nine amino acids, four dipeptides, one tripeptide and several types of proteins were studied by reflectance and absorption FT-IR spectroscopy. Mid-IR and, where necessary, Raman frequencies were applied to assign the combination and overtone bands occurring in the near-IR region. On one occasion, an isotope effect (deuterization) was also employed for the recognition of the combinations involving N+H3 vibrations. The overtone and combination bands of N+H3, CH3, CH2, NH2 and amide groups were, in part, assigned. On the basis of this assignment and with the use of second derivative and/or Fourier deconvolved spectra, attempts were made to recognize the characteristic features of glutamine and asparagine components of some proteins. The IR spectra were taken on an FT-IR spectrometer with a configuration allowing spectra to be measured in the 10,000 to 400 cm−1 range.  相似文献   

9.
We present an approach based on the statistical associating fluids theory (SAFT) to predict the solubility of amino acids in aqueous and aqueous-electrolyte solutions. This approach can describe the association interactions and their effects on the solubility of amino acids. Using the experimental data of activity coefficients of amino acids in water, the parameters of SAFT model for amino acids are obtained. The solubility of several amino acids in the temperature range of 273.15–373.15 K is predicted. Results obtained from the model are in a good accordance with the experimental data. Also, we examine the effect of pH on the solubility of dl-methionine. Addition of an extra amino acid to the binary solution of amino acid + water makes the system more complex. To check the accuracy of model, we study the ternary solution of dl-serine + dl-alanine + water and dl-valine + dl-alanine + water. Predicted results depict that the proposed model has the ability to describe the ternary solution of amino acids, accurately. Finally, the solubility of amino acids in aqueous-electrolyte solutions is investigated. The long-range interactions caused by the presence of ions affects the solubility of amino acids, leading them to be salted in or out. To treat this kind of interaction, the restrictive primitive mean spherical approximation (RP-MSA) is coupled with the SAFT equation of state. The proposed model can accurately predict the solubility of amino acids in aqueous-electrolyte solutions.  相似文献   

10.
Epoxide ring opening driven alkalinization process was explored with the aim of preparing layered double hydroxide (LDH) phases on demand, at room temperature. Employing iodide as nucleophilic agent, the precipitation reaction can be driven under much lower halide concentrations. This scenario favors the selective intercalation of concomitant bulky oxo anions as nitrate or perchlorate in the LDH products, allowing for the one-pot synthesis of an LDH able to delaminate in formamide. Even large dicarboxylic acids, O2C-(CH2)n-CO2, with n up to 8, can be quantitively intercalated within the growing LDH phase, providing a versatile one-pot route for hybrid LDHs as well. Under the mild conditions employed, governed by a continuous pH rise from a starting acid condition, a MII to M*III ratio of 2 prevails, independently from the overall cationic composition. However, after moderate hydrothermal aging LDH phases bearing a cationic ratio higher than 2 could result. The solubility of a given chloride-containing MII2M*III LDH can be approximated as a linear combination of the solubility of the pure hydroxylated phases of the constitutive cations, M(OH)2 and M*(OH)3.  相似文献   

11.
Effect of nucleofugacity of leaving group, X, nucleophilicity/basicity of the formed counteranion, X or MtXn+1, and groups stabilizing positive charge in the carbocation, R+, generated from initiator, RX, as well as metal and ligands in the Lewis acids, MtXn, strength of protonic acids, HA, added salts, NR4+X, nucleophiles, Nu, some other additives and also reaction conditions on controlled/“living” carbocationic polymerization is discussed. The role of all of the components is explained and their rational design for various monomers which can be polymerized cationically to well defined polymers and copolymers is provided.  相似文献   

12.
A biomimetic nickel bis‐diphosphine complex incorporating the amino acid arginine in the outer coordination sphere was immobilized on modified carbon nanotubes (CNTs) through electrostatic interactions. The functionalized redox nanomaterial exhibits reversible electrocatalytic activity for the H2/2 H+ interconversion from pH 0 to 9, with catalytic preference for H2 oxidation at all pH values. The high activity of the complex over a wide pH range allows us to integrate this bio‐inspired nanomaterial either in an enzymatic fuel cell together with a multicopper oxidase at the cathode, or in a proton exchange membrane fuel cell (PEMFC) using Pt/C at the cathode. The Ni‐based PEMFC reaches 14 mW cm−2, only six‐times‐less as compared to full‐Pt conventional PEMFC. The Pt‐free enzyme‐based fuel cell delivers ≈2 mW cm−2, a new efficiency record for a hydrogen biofuel cell with base metal catalysts.  相似文献   

13.
For a better understanding on the functions of DMSO in biological systems at a relatively lower concentration, apparent molar volumes of three typical amino acids, glycine, l-alanine and l-serine in (DMSO + water) mixtures were determined and the transfer volumes from water to the mixtures were evaluated. Together with static light scattering measurement, the results were utilised to reveal the microscopic solvent structure of (DMSO + water) mixtures and its influence on the interaction between DMSO and amino acids from a clustering point of view. The results demonstrate that the interaction between amino acids and DMSO is greatly related to the clustering structure of the mixed solvent and that amino acids interacted with already established solvent clusters. The linear dependence of transfer volume of amino acids on DMSO concentration up to 2.0 mol  dm−3 could be attributed to the increasing interaction with (DMSO)1(H2O)n clusters. The formation of (DMSO)m(H2O)n cluster via hydrophobic aggregating at higher DMSO concentration led to a decrease in hydrophobic effect of DMSO and its hydrophobic–hydrophilic and hydrophobic–hydrophobic interaction with amino acids. The structure change of solvent and the interaction between amino acid residues and DMSO was reflected by the solvation of proteins. It was found that dependence of hydrodynamic radius of bovine serum albumin and lysozyme on DMSO concentration was the same and similar to that of static light scattered by the mixed solvent, regardless of the difference in conformational change between the two proteins.  相似文献   

14.
The solubilities of two cinnamic acid derivatives, namely p-coumaric acid and caffeic acid, in six 1-alkyl-3-methyl imidazolium based ionic liquids composed of the PF6, BF4, TFO and TF2N anions, and in two organic solvents, t-pentanol and ethyl acetate, have been measured at the temperature range of about (303 to 317) K. The p-coumaric acid was found to be more soluble than caffeic acid in all studied solvents. Higher solubilities of both acids were observed in the ionic liquids composed of the BF4 and TFO anions. The increase of the alkyl chain length on the cation invokes a decrease in solubility in the case of hydrophilic ionic liquids composed of BF4 anion, while in the case of hydrophobic ones composed of PF6 anion an increase in the solubility is observed. Between the two organic solvents t-pentanol is better solvent than ethyl acetate for both acids. Moreover, using the van’t Hoff equations the apparent Gibbs energy, enthalpy, and entropy of solution were calculated. Finally, successful correlation of the experimental data was achieved with the UNIQUAC and the NRTL activity coefficient models, while poor predictions of the solubility of the two acids in the organic solvents were obtained with two UNIFAC models.  相似文献   

15.
The rate of oxidation of amino acids (AA) by N-Bromoacetamide (NBA) was studied in aqueous buffered medium at 35°C. The rate of disappearance of [NBA] is catalyzed by the Br produced from the reduction of NBA. Analysis of the autocatalyzed reaction gives the kinetic data for the oxidation of bromide ion by NBA. The results suggest that the protonated NBA reacts with Br to form Br2 which rapidly oxidizes amino acids. The rate constant for the reaction between protonated NBA and Br at 35°C is estimated. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
This paper discusses the alkaline ion (Na+) role in the uphill transport of amino acids through a bulk liquid membrane. The aqueous phases (source phase - S and receiving phase - R) are made up of equimolar concentrations of amino acid (4.38 mM p-aminobenzoic acid (PABA)) and alkaline ion (75 mM Na+). A chloroform solution containing 5 mM dibenzo-18-crown-6 (DB18C6) represents the bulk liquid membrane (M). The data obtained show that at the S-M interphase, the amino acid is coupled with the carrier via the H3N+ group rather than being transported to the R-M interphase, where Na+ substitutes the amino acid. If Na+ is absent, the amino acid is transported to the opposite direction. These results support the hypothesis that the presence of Na+ ion in the aqueous phases assures the ‘biological’ direction of aminobenzoic acids transport through membranes.  相似文献   

17.
Abstract

Previous developed theories were applied in explaining the mechanism for the salting-in and -out of various amino acids. Glycine is salted-in according to the cationic sequences Li+ > Na+ > K+ > Rb+ and Ca2+ > Ba2+ > Sr2+. The ability of a cation to increase the solubility of an amino acid therefore corresponds to the destruction of the ion-ion bond between the - CO-2and the -NH+ 2group of the amino acid by forming an insoluble ion-ion bond between the added cation and the - CO?2 group. This insolubilizing effect produces a positive charge on the amino acid. If, however, the anion of the added salt forms a relatively insoluble ion-ion bond with the -NH+2 group of the amino acid, then the effect is minimized because now both charges on the amino acid are reduced. Consequently, the more insoluble the cation amino acid salt and the more soluble the anion amino acid salt (or vice versa), the greater will be the salting-in effect. Titration of either charged group on the amino acid zwitterion has the same effect, since now the ion-ion bond of the amino acid is again destroyed. Aliphatic and carboxylic acid groups also effect the salting-in sequence, since these groups are salted-out by addition of salt when D± < DH2o. These mechanisms explain how leucine is first salted-out, then salted-in (at 4 M) and finally salted-out again (at 9 M) in LiCl solutions. Urea salts-in hydrophobic amino acids by increasing the dielectric constant and salts-out polar amino acids by increasing the interaction between the two charge groups on the amino acid. Glycine reverses the salting-in effect of NaCl on asparagine by competing for the Na+ ion.  相似文献   

18.
19.
The electrooxidation of HCOONa was carried out over a wide range of pH on Pt. HCOO and its associated form of HCOOH do not show any difference in electrochemical behaviour. A voltammetric study demonstrates the formation of two kinds of poisoning species in the hydrogen (X1) and double-layer (X2) regions. Their dependences on the potential and pH were examined. Constant polarization measurements give the rate expression, i = kH+)−0.43 exp(0.4Fφn/RT), independent of the concentrations of HCOO and HCOOH. The rate-determining step is concluded to be HCOO (a) → COO (a)+H+ + e or HCOOH(a) → COOH(a)+H+ + e. The negative reaction order with respect to H+ was explained through the retarding action of X2. The nature of X1 and X2 is discussed.  相似文献   

20.
《Polyhedron》1987,6(5):1049-1052
Arsenate hydroxyapatite, Ca10(AsO4)6(OH)2 (AsHA), was synthesized by precipitation at 100°C, and characterized by X-ray studies, IR spectra, electron microscopy, and thermography (TGA, DTA and DTG). The solubility of AsHA was investigated between temperatures of 308 and 323 K, and in the pH range 4.0–8.0. It exhibited a retrograde solubility. The solubility behaviour can be explained on the basis of the chemical potential and free energy in solution. Standard thermodynamic parameters, ΔHθ, ΔSθ and ΔCpθ, are reported. The solubility products of AsHA were found to be 4.0 x 10−91, 2.7 x 10−91, 2.1 x 10−91 and 1.2 x 10−91 at temperatures of 3.08, 313, 318 and 323 K, respectively.  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