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.
With the specific aim of calculating the acidity equilibrium constant (Ka) of carboxylic acids in aqueous solution we investigated the solute-solvent interactions of these acids and their corresponding anions. The pKa (−lg Ka) values have been calculated using density functional theory (DFT). The polarized continuum model (PCM) is used to describe the solvent. Using these methods, we successfully predicted the pKas of 66 carboxylic acids in aqueous with the average error of 0.5 in pKa units. Two different thermodynamic cycles have been studied. The theoretical values are in better agreement with the experimental results for those acids with moderate strength of acidity with the pKa value higher than 3.  相似文献   

2.
M. Remko 《Chemical Papers》2007,61(2):133-141
Computational chemical methods have been used to correlate the molecular properties of the 10 ACE inhibitors (captopril, enalapril, perindopril, lisinopril, ramipril, trandolapril, quinapril, fosinopril, benazepril, and cilazapril) and some of their active metabolites (enalaprilat, perindoprilat, ramiprilat, trandolaprilat, quinaprilat, fosinoprilat, benazeprilat, and cilazaprilat). The computed pK a values correlate well with the available experimental values. In the dicarboxylic ACE inhibitors, the carboxyalkyl carboxylate group of the ACE inhibitors studied is more acidic than the C-terminal carboxylate. However, at physiological pH = 7.4 both carboxyl groups of ACE inhibitors are completely ionized and the dicarboxyl-containing ACE inhibitors behave as strong acids. The available experimental partition coefficients of these ACE inhibitors investigated are well reproduced by the neural network-based ALOGPs and the fragment-based KoWWiN methods. All parent drugs (and prodrugs), with the exception of fosinopril, are compounds with low lipophilicity. Calculated pK a, lipophilicity, solubility, absorption, and polar surface area of the most effective ACE inhibitors for the prevention of myocardial infarction, perindopril and ramipril, were found similar. Therefore, it is probable that the experimentally observed differences in the survival benefits in the first year after acute myocardial infarction in patients 65 years of age or older correlate closely to the physicochemical and pharmacokinetic characteristics of the specific ACE inhibitor that is used.  相似文献   

3.
A new and fast method to determine acidity constants of monoprotic weak acids and bases by capillary zone electrophoresis based on the use of an internal standard (compound of similar nature and acidity constant as the analyte) has been developed. This method requires only two electrophoretic runs for the determination of an acidity constant: a first one at a pH where both analyte and internal standard are totally ionized, and a second one at another pH where both are partially ionized. Furthermore, the method is not pH dependent, so an accurate measure of the pH of the buffer solutions is not needed. The acidity constants of several phenols and amines have been measured using internal standards of known pKa, obtaining a mean deviation of 0.05 pH units compared to the literature values.  相似文献   

4.
Jia-Ning Li  Yao Fu 《Tetrahedron》2006,62(18):4453-4462
A first-principle theoretical protocol was developed, which could successfully predict the pKa values of a number of amines and thiols in DMSO with a precision of about 1.1 pKa unit. Using this protocol we calculated the pKa values of diverse types of organophosphorus compounds in DMSO. The accuracy of these predicted values was estimated to be about 1.1 pKa because phosphorus is in the same group as nitrogen and in the same period as sulfur. The theoretical predictions were also consistent with all the available experimental data. Thus, a scale of reliable pKa values was constructed for the first time for organophosphorus. These pKa values would be helpful to synthetic chemists who need to design the experimental conditions for handling deprotonated organophosphorus. On the basis of these pKa values we also studied, for the first time, some interesting topics such as the substituent effects on the pKa values of various types of organophosphorus, and the differences between the pKa values of organophosphorus and organic amines.  相似文献   

5.
In this work, we first studied the pH-dependent characteristic of chromenoquinoline. Based on this, we then designed and synthesized two novel chromenoquinoline derivatives that can act as fluorescent pH sensors. The pKa values of two novel chromenoquinoline derivatives can be modulated from 2.32 to 4.38 and 6.27 by introducing EDG on the backbone of chromenoquinoline. Furthermore, we demonstrate that the sensor 4 can be used as a ratiometric fluorescent pH sensor for fluorescence imaging in living cells.  相似文献   

6.
In this study, pKa values were determined using the dependence of the retention factor on the pH of the mobile phase for three ionizable substances, namely, enalapril, lercanidipine and ramipril (IS). The effect of the mobile phase composition on the ionization constant was studied by measuring the pKa at different methanol-water mixtures, ranging between 50 and 65% (v/v), using LC-DAD method. Two simple, accurate, precise and fully validated analytical methods for the simultaneous determination of enalapril and lercanidipine in combined dosage forms have been developed. Separation was performed on an X-Terra RP-18 column (250 mm × 4.60 mm ID × 5 μm) at 40 °C with the mobile phase of methanol-water 55:45 (v/v) adjusted to pH 2.7 with 15 mM orthophosphoric acid. Isocratic elution was performed in less than 12 min with a flow rate of 1.2 mL min−1. Good sensitivity for the analytes was observed with DAD detection. The LC method allowed quantitation over the 0.50-20.00 μg mL−1 range for enalapril and lercanidipine. The second method depends on first derivative of the ratio-spectra by measurements of the amplitudes at 219.7 nm for enalapril and 233.0 nm for lercanidipine. Calibration graphs were established for 1-20 μg mL−1 for enalapril and 1-16 μg mL−1 lercanidipine, using first derivative of the ratio spectrophotometric method. Both methods have been extensively validated. These methods allow a number of cost and time saving benefits. The described methods can be readily utilized for analysis of pharmaceutical formulations. The methods have been applied, without any interference from excipients, for the simultaneous determination of these compounds in tablets. There was no significant difference between the performance of the proposed methods regarding the mean values and standard deviations.  相似文献   

7.
Disubstituted peptide ferrocenes conjugates were prepared from ferrocenedicarboxylic acid (Fc(OH)2) and glycineethylester. After conversion of the resulting ester Fc(Gly-OEt)21 into the corresponding acid Fc(Gly-OH)22 by ester hydrolysis, significant structural changes take place in the way the molecules interact with each other. Complex 1 adopts a 1,3-conformation showing extensive intermolecular H-bonding forming 1-D chains, whereas complex 2 displays a compact 1,2-conformation in which the NH on one strand engage in strong intramolecular cross-strand H-bonding involving the amide CO on the opposite strand. Additional intermolecular H-bonding in 2 allows the formation of a 2-D net. In essence, ester-deprotection allows us to switch the ferrocene conformation and the H-bonding pattern.  相似文献   

8.
A thermodynamic cycle to calculate pKa values (Minus log of acid dissociation constants) of hydroxamic acids is presented. Hydroxamic acids exist mainly as amide isomers in the aqueous medium. The amide form of hydroxamic acids has two deprotonation sites and may yield either an N-ion or an O-ion upon deprotonation. The thermodynamic cycle proposed includes the gas-phase N–H deprotonation of the hydroxamic acid, the solvent phase transformation of the N-ion to the O-ion and the solvation of the hydroxamic acid molecule and the O-ion in water. The CBS-QB3 method was employed to obtain gas-phase free energy differences between 12 hydroxamic acids and their respective anions. The aqueous solvation Gibbs free energy changes were calculated at the HF/6-31G(d)/CPCM and HF/6-31+G(d)/CPCM levels of theory using HF/6-31+G(d)/CPCM geometries. For the proton, literature values of the gas-phase free energy of formation and the solvation free energy change were used. The free energy change for the transformation of the N-ion to O-ion in the aqueous medium was calculated by employing CBS-QB3/CPCM in the aqueous medium. For this, the hydroxamic acids were divided in two classes according to the substituent at the carbonyl carbon. A common transformation free energy difference for aliphatic substituted hydroxamic acids and a separate common transformation free energy difference for aromatic substituted hydroxamic acids were obtained. The pKa calculation yielded a root mean square error of 0.32 pKa units.  相似文献   

9.
Carmine Gaeta  Placido Neri 《Tetrahedron》2008,64(22):5370-5378
Water-soluble p-sulfonatocalix[7]arene 1 has been synthesized in good yield through standard procedures and its conformational preferences have been investigated by Monte Carlo conformational searches. The acid-base properties of 1 were investigated by means of potentiometric titration, obtaining pKa values in agreement with those reported for other p-sulfonatocalix[n]arene homologs. The binding ability of 1 toward organic quaternary ammonium cations such as Diquat (2), Paraquat (3), and Chlormequat (4) was investigated by means of 1H NMR titrations in D2O at pD=7.3, DOSY NMR measurements, and 2D ROESY NMR spectroscopy. Spectrofluorimetry proved to be a useful method for the determination of trace amounts of 2 and 3 in aqueous solution by using Acridine Orange bound to 1 as a chemical indicator.  相似文献   

10.
This article describes the synthesis and in vitro biological affinities of (poly)fluorinated neprilysin inhibitors. Two series of inhibitors with F-substitution of the central benzimidazole platform of the ligands and the benzylic vector to fill the S1’ pocket of NEP were investigated. The S1’ pocket was found to be highly fluorophobic, and F-substitution led to significantly decreased binding affinities of inhibitors. This result is explained by electrostatically unfavorable close contacts of organic fluorine with the negatively polarized π-surfaces of surrounding aromatic amino acid side chains. In contrast, the protein environment around the benzimidazole platform, with three electropositive guanidinium side chains of Arg residues, was found to provide a fluorophilic environment. Overall, the data support that organic fluorine, with its high negative charge density prefers to orient into electropositive regions of receptor sites. pKa measurements of fluorinated ligands provided several simple patterns for the prediction of pKa values of benzimidazoles, important building blocks in medicinal chemistry.  相似文献   

11.
In this paper the validation of pKa determination in MDM-water mixtures is presented. The MDM-water mixture is a new multicomponent cosolvent mixture (consisting of equal volumes of methanol, dioxane and acetonitrile, as organic solvents) that dissolves a wide range of poorly water-soluble compounds. The cosolvent dissociation constants (psKa) of 50 chemically diverse compounds (acids, bases and ampholytes) were measured in 15-56 wt% MDM-water mixtures by potentiometric or spectrophotometric titration and the aqueous pKa values obtained by extrapolation. Three different extrapolation procedures were compared in order to choose the best extrapolation in MDM-water mixture using a sub-set of 30 water-soluble compounds. The extrapolated results are in good agreement with pKa values measured in aqueous medium. No significant difference was found among these extrapolation procedures thus the widely used Yasuda-Shedlovsky plot was proposed for MDM cosolvent also. Further we also present that the single point estimation based on measurement in 20%/v MDM-mixture using a general calibration equation may be suitable for rapid pKa determination in the early phase of drug research.  相似文献   

12.
The pKa values of a series of fluoroalkanesulfonylamides were measured by potentiometric titration. Different kinds of alkyl halides and tosylates were employed to investigate the nucleophilicity of fluoroalkanesulfonylamides. Fluoroalkanesulfonylamides with longer fluoroalkyl chain have weaker nucleophilicity.  相似文献   

13.
The chemical kinetics, studied by UV/Vis, IR and NMR, of the oxidative addition of iodomethane to [Rh((C6H5)COCHCOR)(CO)(PPh3)], with R = (CH2)nCH3, n = 1-3, consists of three consecutive reaction steps that involves isomers of two distinctly different classes of RhIII-alkyl and two distinctly different classes of RhIII-acyl species. Kinetic studies on the first oxidative addition step of [Rh((C6H5)COCHCOR)(CO)(PPh3)] + CH3I to form [Rh((C6H5)COCHCOR)(CH3)(CO)(PPh3)(I)] revealed a second order oxidative addition rate constant approximately 500-600 times faster than that observed for the Monsanto catalyst [Rh(CO)2I2]. The reaction rate of the first oxidative addition step in chloroform was not influenced by the increasing alkyl chain length of the R group on the β-diketonato ligand: k1 = 0.0333 ([Rh((C6H5)COCHCO(CH2CH3))(CO)(PPh3)]), 0.0437 ([Rh((C6H5)COCHCO(CH2CH2CH3))(CO)(PPh3)]) and 0.0354 dmmol−1 s−1 ([Rh((C6H5)COCHCO(CH2CH2CH2CH3))(CO)(PPh3)]). The pKa and keto-enol equilibrium constant, Kc, of the β-diketones (C6H5)COCH2COR, along with apparent group electronegativities, χR of the R group of the β-diketones (C6H5)COCH2COR, give a measurement of the electron donating character of the coordinating β-diketonato ligand: (R, pKa, Kc, χR) = (CH3, 8.70, 12.1, 2.34), (CH2CH3, 9.33, 8.2, 2.31), (CH2CH2CH3, 9.23, 11.5, 2.41) and (CH2CH2CH2CH3, 9.33, 11.6, 2.22).  相似文献   

14.
The difficulties in estimating uncertainty of pKa values determined in nonaqueous media are reviewed and two different uncertainty estimation approaches are presented and applied to the pKa values of the compounds on a previously established self-consistent spectrophotometric basicity scale in acetonitrile. One approach is based on the ISO GUM methodology (the “ISO GUM” approach) and involves careful analysis of the uncertainty sources and quantifying the respective uncertainty components. The second approach is based on the standard-deviation-like statistical parameter that has been used for characterization of the consistency of the scale (the “statistical” approach). It is demonstrated that the ISO GUM approach somewhat overestimates the uncertainty. The statistical approach is based on long-term within-laboratory statistical data and it is demonstrated that it underestimates the uncertainty. In particular it neglects the laboratory bias effects that are taken into account at least to some extent by the ISO GUM approach. Thus, together these two approaches allow to “bracket” the uncertainties of the pKa values on the scale. The uncertainties of the pKa values are defined in two different ways. Definition (a) includes the uncertainty of the pKa of the reference base (anchor base of the scale) pyridine. Definition (b) excludes it. It is demonstrated that both definitions have their virtues. Definition (a) leads to the uncertainty ranges of 0.12-0.22 and 0.12-0.14 pKa units at standard uncertainty level for different bases using the ISO GUM and statistical approach, respectively. Definition (b) leads to the uncertainty ranges of 0.04-0.19 and 0.02-0.08 pKa units, respectively. The uncertainty of the pKa of a given base is dependent on the quality of the measurements involved and on the distance from the reference base on the scale. The importance of the correlation between the pKa values of bases belonging to the same scale is stressed.  相似文献   

15.
Apparent pK values of the wine pigment, 5-carboxypyranomalvidin-3-glucoside (vitisin A), were determined using UV-vis spectroscopy, viz. pKa1=0.98 (±0.10), pKH1=4.51 (±0.03) and pKH2=7.57 (±0.02). An additional ionisation constant at high pH (pKa4=8.84±0.06) was established by high-voltage paper electrophoresis. These data in conjunction with previously published pKa values determined by high-voltage electrophoresis suggest that in wine (pH 3.2-3.8), 5-carboxypyranomalvidin-3-glucoside exists as a complex mixture of hydrated and non-hydrated, partially ionised species with the predominant species being the quinonoidal base (λmax 498 nm).  相似文献   

16.
A new method of the European Pharmacopoeia (Ph. Eur.)—alkalimetry in alcohol-water mixture with potentiometric end-point detection—for the assay of halide salts of alkaloids and other organic N-bases has been investigated using 13 substances. The results were compared to those obtained by nonaqueous direct acidimetry. In general, our measurements show that the Ph. Eur. method may be regarded as an environment-friendly, precise, reproducible approach, with average S.D. ±0.33. However, in six cases out of 13, the method did not work. In these cases, the first inflexion point, which should appear due to the neutralization of a defined, small volume of HCl added to the solution of the sample before the titration, was missing on the potentiometric titration curves. This was observed for papaverine chloride, pilocarpine chloride, pyridoxine chloride, thiamine chloride, histamine dichloride and noscapine chloride; this missing inflexion point hampered the exact measurement of the alkaloid content. This study shows that the method, in the present official form, can be applied only for compounds with pKa values of seven or higher. For the salts of weaker bases, a modified approach (titration in 70% ethanol, without addition of HCl) is proposed.  相似文献   

17.
18.
The synthesis of a new pigment with a bluish color was obtained from the reaction of methylpyranomalvidin-3-glucoside with sinapaldehyde and its formation mechanism seems to involve a charge-transfer reaction pathway. The structure of this compound was fully characterized by LC/DAD-MS and NMR. Its equilibrium forms present in water at different pH values and the respective ionization constants were determined by UV-visible spectroscopy. The results revealed the presence of three equilibria involving only deprotonation reactions at the 7-OH and 4′-OH positions of the pyranoanthocyanin moiety and at the 4′″-OH of the syringol moiety (pKa1 = 3.64 ± 0.01; pKa2 = 8.02 ± 0.01; pKa3 = 11.19 ± 0.01).  相似文献   

19.
The quenching of the anionic dye 8-hydroxypyrene-1,3,6-trisulfonic acid trisodium salt (pyranine) with three different boronic acid-substituted benzyl viologens was determined, and the fluorescence signal modulation obtained upon addition of glucose to the dye/quencher system was also studied. The benzyl viologen that contains boronic acids in the ortho-position (o-BBV) was found to display unique behavior, which can be rationalized by a charge neutralization mechanism facilitated by an intramolecular interaction between sp3 boronate and the quaternary nitrogen of the viologen. Potentiometric titration and 11B NMR spectroscopy were used to generate pH profiles for the boronic acids, which provide additional evidence for the proposed mechanism.  相似文献   

20.
A fast method for the determination of acidity constants by CZE has been recently developed. This method is based on the use of an internal standard of pK(a) similar to that of the analyte. In this paper we establish the reference pK(a) values of a set of 24 monoprotic neutral acids of varied structure that we propose as internal standards. These compounds cover the most usual working pH range in CZE and facilitate the selection of adequate internal standards for a given determination. The reference pK(a) values of the acids have been established by the own internal standard method, i.e. from the mobility differences between different acids of similar pK(a) in the same pH buffers. The determined pK(a) values have been contrasted to the literature pK(a) values and confirmed by determination of the pK(a) values of some acids of the set by the classical CE method. Some systematic deviations of mobilities have been observed in NaOH buffer in reference to the other used buffers, overcoming the use of NaOH in the classical CE method. However, the deviations affect in a similar degree to the test compounds and internal standards allowing thus, the use of NaOH buffer in the internal standard method. This fact demonstrates the better performance of the internal standard method over the classical method to correct mobility deviations, which together with its fastness makes it an interesting method for the routine determination of accurate pK(a) values of new pharmaceutical drugs and drug precursors.  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