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.
Budevsky O 《Talanta》1989,36(12):1209-1216
A new approach for selection of a suitable solvent system as a medium for non-aqueous acid—base titration is proposed. The essence of the approach is the development of a new criterion called “effectivity”. The latter is based on consequences of the Brønsted and Izmailov acid—base theories and represents a quantitative measure for improving or worsening the titration conditions of acids and bases in non-aqueous solvents as compared with water. The “effectivity” E is given by the relation E = ΔpKs - ΔpKs where ΔpKs is the difference between the logarithmic values of the autoprotolysis constants of water and the solvent in question, and ΔpK is the so-called medium effect. The latter is a constant value which shows that acids and bases with the same charge alter their strength to the same extent when transferred from water into a non-aqueous solvent. The medium effect is calculated by statistical treatment of a great number of acid—base constants determined experimentally both in water and the non-aqueous solvent in question. The effectivity of the solvents most often used in non-aqueous acid—base titrimetry, determined by this approach, shows that in many cases these solvents offer significant advantages over water, but drawbacks are also observed. Some limitations of the approach are discussed. Special attention is paid to dimethylsulphoxide and its mixtures with water, which prove to be highly effective media for the acid—base titration of many substances.  相似文献   

2.
Potentiometric titration has been conducted to systematically examine the acid–base properties of the cell surfaces of Escherichia coli K-12 and Bacillus brevis as a function of growth phase, nitrogen source (ammonium or nitrate), and carbon to nitrogen (C:N) ratio of the growth substrate. The two bacterial species revealed four distinct proton binding sites, with pKa values in the range of 3.08–4.05 (pK1), 4.62–5.57 (pK2), 6.47–7.30 (pK3), and 9.68–10.89 (pK4) corresponding to phosphoric/carboxylic, carboxylic, phosphoric, and hydroxyl/amine groups, respectively. Two general observations in the data are that for B. brevis the first site concentration (N1), corresponding to phosphoric/carboxylic groups (pK1), varied as a function of nitrogen source, while for E. coli the fourth site concentration (N4), corresponding to hydroxyl/amine groups (pK4), varied as a function of C:N ratio. Correspondingly, it was found that N1 was the highest of the four site concentrations for B. brevis and N4 was the highest for E. coli. The concentrations of the remaining sites showed little variation. Finally, comparison between the titration data and a number of cell surface compositional studies in the literature indicates one distinct difference between the two bacteria is that pK4 of the Gram-negative E. coli can be attributed to hydroxyl groups while that of the Gram-positive B. brevis can be attributed to amine groups.  相似文献   

3.
Granular nanoporous activated carbon prepared from polyacrylonitrile (PAN) was investigated as a means of removing weak aromatic acids from aqueous solutions. This carbon is highly nanoporous, the BET surface area being 544 m2/g with Vtot=0.278 cm3/g, and Vmicro=0.266 cm3/g. Aqueous treatment reduced the surface area to 364 m2/g.

Granular nanoporous carbon prepared from PAN contains O and N related surface functional groups. The surface concentration of both oxygen and nitrogen atoms was found by XPS to be 5.3%. Surface groups containing these hetero atoms are responsible for the acid–base character of this carbon in aqueous solutions. The pHPZC of the carbon is 8.4. The microporous pore network produces a wide hysteresis loop, observed when the granular carbon was studied by continuous titration. This loop, which was found to diminish, but not to disappear completely when the particle size was reduced, is attributed to irreversible hydrolysis of surface esters and/or lactones. The surface concentration of the functional groups titrated by the equilibrium Boehm method is 449.2 μeq./g, of which 112.9 μeq./g have acidic and 336.3 μeq./g have basic character. Within the acidic species three subgroups, namely the regions pKa<6.37, 6.37Ka<10.25 and 10.25Ka<15.74.

The acid–base behavior influences the adsorption performance of the granular carbon in aqueous solutions of weak aromatic acids, in this case phenol and 2,3,4-trichlorophenol. Both the adsorption capacity and the overall interaction parameter, K (both derived from a fit to the Langmuir equation) depend on the adsorbed species and on the pH. The former is a consequence of the different water solubilities of the solute molecules, while the latter stems from the pH sensitivity of both the surface functional groups and these weak acids. The K values show a sequence pH=3K values indicate different adsorption mechanisms. The effect of pH is more marked for 2,3,4-trichlorophenol.  相似文献   


4.
Barbosa J  Roses M  Sanz-Nebot V 《Talanta》1988,35(12):1013-1018
A series of acid—base indicators in acetonitrile has been studied. Their chromaticity coordinates, pKa values, transition pH ranges, pH of maximum colour change, optimum concentration for titration, and quality of colour change in acetonitrile have been determined, together with the effect of ionic strength on their properties. Various bases and their mixtures have been titrated to test the practical usefulness of the indicators.  相似文献   

5.
Though the H-bond is well characterized as a D–H:A three-center-four-electron interaction, the formulation of a general H-bond theory has turned out to be a rather formidable problem because of the extreme variability of the bonds formed (for instance, O–HO energies range from 0.1 to 31 kcal mol−1). This paper surveys our previous contributions to the problem, including: (a) the H-bond chemical leitmotifs (CLs), showing that there are only four classes of strong H-bonds and one of moderately strong ones; (b) the PA/pKa equalization principle, showing that the four CLs forming strong H-bonds are actually molecular devices apt to equalize the acid–base properties (PA or pKa) of the H-bond donor and acceptor groups; (c) the driving variable of the H-bond strength, which remains so identified as the difference ΔpKa=pKAH(D–H)−pKBH(A–H+) or, alternatively, ΔPA=PA(D)−PA(A); and, in particular, (d) the transition-state H-bond theory (TSHBT), which interprets the H-bond as a stationary point along the complete proton transfer pathway going from D–HA to DH–A via the DHA transition state. TSHBT is verified in connection with a series of seven 1-(X-phenylazo)-2-naphthols, a class of compounds forming a strong intramolecular resonance-assisted H-bond (RAHB), which is switched from N–HO to NH–O by the decreasing electron-withdrawing properties of the substituent X. The system is studied in terms of: (i) variable-temperature X-ray crystallography; (ii) DFT emulation of stationary points and full PT pathways; (iii) Marcus rate-equilibrium analysis correlated with substituent LFER Hammett parameters.  相似文献   

6.
Fan J  Wang J  Ye C 《Talanta》1998,46(6):1285-1292
The acid dissociation constants (Ka), base dissociation constants (Kb) and the autoprotolysis constants (Ks) for 2,2′-bipyridyl in water and in water+alcohol(methanol, ethanol, iso-propanol) mixed solvents have been determined at 25°C and an ionic strength of 0.1 mol l−1, from a direct potentiometric method based on the treatment of the data of a single pH titration. It has been shown that Ka increases, whereas Kb and Ks decrease, with increasing proportion of the alcohol in the mixed solvents. Linear relations between pKa, pKb, pKs and the mole fraction of the alcohol were observed in the composition range investigated. These results are discussed in terms of the properties of solvent and the interactions of the different species existing in dissociation equilibrium with solvents. It is concluded that the higher stabilization of both 2,2′-bipyridyl and its protonated form by dispersion forces and of the proton by its interaction with solvent molecules in the mixed solvents compared with that in water are largely responsible for the observed changes of pKa with composition. On the other hand, the low stabilization of OH in the mixed solvents relative to that in water and the electrostatic effect are the main factors in determining the solvent effect on pKb.  相似文献   

7.
Xia YX  Chen JF  Choppin GR 《Talanta》1996,43(12):2073-2081
Equilibria in the system of Nd(III) and Th(IV) with 8-hydroxyquinoline (oxine), thenoyltrifluoroacetone (HTTA) and 1,10-phenanthroline (phen) in 5.0 m NaCl solution have been investigated by spectroscopy and potentiometry. The solubility and deprotonation constants of the three organics were measured to be: pKs = 3.09 ± 0.01, pKa1 = 5.82 ±0.02, pKa2= 10.00 ±0.01 for oxine; pKs = 2.49 ± 0.01, pKa1 = 6.47 ±0.03 for HTTA; pKs = 2.86 ± 0.02, pKa2 = 5.82 ± 0.05 for phen. The stabilities of the corresponding metal complexes are in the order M(oxine) > M(TTA) > M(phen), where M = Nd(III), Th(IV). For all three organic ligands, the Th(IV) complexation is stronger than that of Nd(III).  相似文献   

8.
Li QG  Xu JG  Huang XZ  Chen GZ 《Talanta》1994,41(12):2049-2054
The effects of media properties including buffers, acidity, solvents and surfactant on horseradish peroxidase-catalyzed fluorogenic reaction were investigated. The results showed that the so-called non-fluorescent hydrogen donors were in fact fluorescent. There existed an acid—base equilibrium in the fluorescent dimer product. For p -hydroxyphenylpropionic acid, a pKa value of 8.0 for the product was obtained from its titration curve. The product fluorescence increased with higher pH, however, a longer time was needed to reach the reaction equilibrium due to the pH mismatch problem. Cationic micelles cetyltrimethylammonium bromide and cetyltrimethylammonium chloride could reduce the pH mismatch and offered a way to further increase the determination sensitivity. Finally, a micelle-enhanced flow-injection analysis of horseradish peroxidase is suggested.  相似文献   

9.
The red fluorescent protein Rtms5H146S displays a transition from blue (absorbance λmax 590 nm) to yellow (absorbance λmax 453 nm) upon titration to low pH. The pKa of the reaction depends on the concentration of halide, offering promise for new expressible halide sensors. The protonation state involved in the low pH form of the chromophore remains, however, ambiguous. We report calculated excitation energies of different protonation states of an RFP chromophore model. These suggest that the relevant titration site is the phenoxy moiety of the chromophore, and the relevant base and conjugate acid are anionic and neutral chromophore species, respectively.  相似文献   

10.
Huitink GM  Poe DP  Diehl H 《Talanta》1974,21(12):1221-1229
A satisfactory method for the preparation of Calcein Blue has been devised. Elemental analysis, equivalent weight by neutralization, and the NMR spectrum show the compound to be 4-methylumbelliferone-8-methyleneiminodiacetic acid·0·25H2O. The ultraviolet absorbance and fluorescence have been studied as a function of pH and, combined with potentiometric titration and solubility date, have yielded for the acid dissociation constants the values pK1 = 3·0, pK2 = 6·9, and pK3 = 11·3. These acid functions are identified respectively as carboxyl, phenol, and ammonium ion, the free Calcein Blue being a zwitter-ion. Calcein Blue fluoresces in both acidic and basic solution when excited at a suitable wavelength. The fluorescence of the doubly-charged anion formed on the neutralization of the phenol group, when excited at 360 nm, reaches a maximum at pH 9, and decreases to zero with the neutralization of the ammonium ion; the wavelength of maximum emission is 455 nm. In the presence of calcium, the fluorescence increases with alkalinity up to pH 9 and then remains constant. The calcium derivative is a 1:1 compound, formation constant 107·1. The fluorescence of Calcein Blue at all pH values is quenched by copper(II). The calcium derivative is changed on standing in highly alkaline solution, presumably by ring opening, to another fluorescent material; thus Calcein Blue, although satisfactory as an indicator, is not useful for the direct fluorometric determination of calcium.  相似文献   

11.
Autoprotolysis constants of acetonitrile, propionitrile, nitromethane, ethylene carbonate and dimethyl sulphoxide were determined using a coulometric — potentiometric method with a hydrogen/palladium electrode as generator. The method is based on the titration of a strong base, tetrabutylammonium hydroxide, with H+ ions generated by anodic oxidation of hydrogen dissolved in palladium. The titration was carried out in a galvanic cell with glass and calomel electrodes at 25°C. The pKs values for the investigated solvents are: acetonitrile, 28.8; propionitrile, 24.6; nitromethane, 23.7; ethylene carbonate. 21.5; and dimethyl sulphoxide 29.1. These data are in accordance with those reported in the literature.  相似文献   

12.
The structural evolution of Y zeolite (Si/Al 2.17) weakly dealuminated by hexafluorosilicate (Si/Al 3.13), denoted YD, and exchanged with calcium (CaYD), has been studied after acid–base treatments at 80 °C close to the cation exchange conditions. The stability of the samples was followed by X-ray diffraction and solid-state NMR of 29Si and 27Al; YD zeolite was completely destroyed by treatment with acid pH 2.5 and suffered serious degradation on treatment with alkali at pH 11.8. The introduction of calcium improved the stability of the zeolite in acid and base. In acid CaYD was not destroyed until pH 1. At pH 2, silicon and aluminium were extracted and an amorphous phase was formed. Base treatment at pH 13 did not affect the calcium-exchanged zeolite.  相似文献   

13.
The theoretical analysis on the zero point of charge (ZPC) and charge density of colloidal particle possessing permanent charges indicates that ZPC determined directly by means of potentiometric titration (PT) should be zero point of net charge (ZPNC) and the permanent charge density (σP) can be obtained from the adsorption amount of H+ and OH-OH-H+) at ZPNC. ZPNC does not change with the electrolyte concentration while the zero point of variable charge (ZPVC) changes with the electrolyte concentration. When σP is zero, ZPNC equals to ZPVC, and only under this condition is ZPC measured directly by PT equal to ZPVC. The relationship between ZPNC, ZNVC, σP, the variable charge density (σV)5 or the total net surface charge density (σT) with pH or electrolyte concentration is derived.  相似文献   

14.
Photo-oxidations of environmental organics in illuminated TiO2 dispersions have implicated surface-bound OH radicals and/or valence band holes. To explore the implications of the former oxidizing entity, six isomeric xylenols (dimethylphenols) were examined by pulsed (nanoseconds to milliseconds) radiolysis methods. The spectral and kinetic characteristics of formation and decay of the transients formed by the reaction of N3, OH and H radicals with these xylenols were assessed in buffered (pH 4, 10−3 M phosphate) aqueous media, where the xylenols exist in their protonated form (pK ≈ 10.19–10.65). The products from the reaction of N3 with 2,6- and 3,4-xylenol were exclusively the corresponding dimethylphenoxyl radicals, formed via electron transfer followed by deprotonation. In contrast, except with 3,4-xylenol, the principal radical intermediates formed initially upon reaction with OH were the corresponding OH adducts, the dihydroxydimethylcyclohexadienyl radicals. 3,4-Xylenol was examined in the pH range 4–10. At pH 8 the initial OH adduct (dihydroxy-3,4-dimethylcyclohexadienyl radical) was subsequently transformed (about 20%–40%) via water elimination into the dimethylphenoxyl radical. In contrast, at pH 9 and 10 the OH adduct and the dimethylphenoxyl radical were formed concurrently (about 60% OH adduct and about 40% dimethylphenoxyl species), the latter through an inner-sphere electron transfer pathway. The switch in behaviour from pH 8 to pH 9 suggests that the pKa of the dihydroxy-3,4-dimethylcyclohexadienyl radical is about 8–9, about 2 pK units below the pKa of the parent substrate (10.4). A mechanism for the conversion of the OH adduct to the dimethylphenoxyl radical is proposed. Reaction of 2,6-xylenol with H radicals gave exclusively the H adduct (hydroxycyclohexadienyl radical), whose spectral characteristics are similar to those of the related OH adduct.  相似文献   

15.
Esteves da Silva JC  Machado AA 《Talanta》1994,41(12):2095-2104
The acid properties of a soil fulvic acid (sfua) were characterized by potentiometric titration with tetrabutylammonium hydroxide in two non-aqueous solvents with high acid—base resolution power N,N-dimethylformamide (DMF) and acetonitrile. Synchronous fluorescence spectroscopy (SyF) was also used to monitor directly the sfua status during the potentiometric titration in DMF. The potentiometric titration curves showed no clear end-point and the analysis of the sets of spectra obtained at increasing neutralization degree, with a self-modeling curve resolution method (SIMPLISMA), revealed the existence of two components with featureless concentration profiles. Internal standards (maleic, salicylic and p-hydroxylbenzoic acids) were used to determine the amounts of acid groups with different acid strengths in the two non-aqueous solvents. It was shown that the variations observed in the SyF spectra sets of the internal standards are not correlated with those observed in the sfua data. The splitting of the sfua groups in the non-aqueous titration curves seems to be forced artificially depending on the standards used.  相似文献   

16.
Zuberbühler AD  Kaden TA 《Talanta》1979,26(12):1111-1118
A fully automatic system for combined spectrophotometric and pH titrations was described in Part I. Its performance in the titration of weak acids and metal complexes is discussed, along with a computer program for numerical treatment of the data, based on Marquardt's modification of the Newton—Gauss non-linear least-squares method. The deprotonation of p-nitrophenol at concentrations of 4 × 10−5 and 4 × 10−6M was studied in order to test the sensitivity. Results identical within the reproducibility of the pH-meter were obtained: pKH = 7.00 ± 0.01 and 7.02 ± 0.01, respectively. Three complexation reactions were studied: (1) the interaction of SCN with the Co2+ complex of 1,4,8,11-tetramethyl-1,4,8,11-tetra-azacyclotetradecane (TMC); five independent experiments gave pK [CoTMC (SCN)+ CoTMC2+ + SCN] = 3.099 ± 0.003: (2) the deprotonation of the Cu2+ complex of 3,7-diazanonanediamide (DANA); five experiments gave pK (CuDANA2+ CuDANAH+−1 + H+) = 7.14 ± 0.01 and pK (CuDANAH+−1 CuDANAH−2 + H+) = 8.38 ± 0.01: (3) for the reaction of Cu2+ with 1,3,7-triazacyclodecane (L), data from different ligand: metal ratios had to be combined to obtain pK (CuL2+ Cu2+ + L) = 16.19 ± 0.01, pK (CuL2+2 CuL2+ + L) = 10.30 ± 0.01, and pK (Cu2L2 (OH)2+2 2 CuL2+ + 2 OH) = 14.58 ± 0.03. Titration curves with a total change in absorbance of as little as 0.03 units could be analysed satisfactorily, extending considerably the useful range of concentrations for spectrophotometric titrations. In combined spectrophotometric/pH titrations the accuracy of the glass electrode is normally the limiting factor. Other equilibrium constants can easily be reproduced with standard errors of less than 0.01 log unit.  相似文献   

17.
A complete electrochemical study and a novel electroanalytical procedure for bromhexine quantitation are described. Bromhexine in methanol/0.1 mol L−1 Britton–Robinson buffer solution (2.5/97.5) shows an anodic response on glassy carbon electrode between pH 2 and 7.5. By DPV and CV, both peak potential and current peak values were pH-dependent in all the pH range studied. A break at pH 5.5 in EP versus pH plot revealing a protonation–deprotonation (pKa) equilibrium of bromhexine was observed. Spectrophotometrically, an apparent pKa value of 4.3 was also determined.

An electrodic mechanism involving the oxidation of bromhexine via two-electrons and two-protons was proposed. Controlled potential electrolysis followed by HPLC–UV and GC–MS permitted the identification of three oxidation products: N-methylcyclohexanamine, 2-amino-3,5-dibromobenzaldehyde and 2,4,8,10-tetrabromo dibenzo[b,f][1,5] diazocine.

DPV at pH 2 was selected as optimal pH for analytical purposes. Repeatability, reproducibility and selectivity parameters were adequate to quantify bromhexine in pharmaceutical forms. The recovery was 94.50 ± 2.03% and the detection and quantitation limits were 1.4 × 10−5 and 1.6 × 10−5 mol L−1, respectively. Furthermore, the DPV method was applied successfully to individual tablet assay in order to verify the uniformity content of bromhexine. No special treatment of sample were required due to excipients do not interfered with the analytical signal. Finally the method was not time-consuming and less expensive than the HPLC one.  相似文献   


18.
Sabry SM  Khamis EF 《Talanta》2000,51(6):159-1231
H-point standard additions method (HPSAM), based on spectrophotometric and spectrofluorimetric measurements, was proposed for simultaneous determination of glafenine (G) and glafenic acid (GA). A study of the absorption spectra of G and GA in various pH media has been carried out. Reasonably resolved UV-absorption spectra were obtained with a solution adjusted at pH 4.5 with citric acid–phosphate buffer. Additionally, the fluorescence properties in aqueous micellar systems of anionic, cationic and non-ionic surfactants were investigated. Well resolved fluorescence spectra were established in aqueous Triton X-100 solution at pH 7.8 (citric acid–phosphate buffer). As a comparative method, UV-derivative spectrophotometry (based on zero-crossing technique) was suggested. First-derivative value at 352 nm (1D352) and second-derivative value at 366 nm (2D366) were selected for the quantification of G and GA, respectively. The relative standard deviations of the proposed methods approximate 2%. The proposed methods were evaluated through the analysis of commercial tablets. The results were accurate and precise.  相似文献   

19.
In the present work we used quantum mechanics calculations to predict the two pKa’s of 5-hydrotryptamine (5-HT). Proton dissociation reaction succeeded to predict the experimental pKa1 corresponding to ionization of the protonated amine group but failed for pKa2 corresponding to ionization of the 5-hydroxyl group. For pKa2, a cluster-continuum model including three water molecules in the first hydration shell around 5-hydroxyl and 5-hydroxide groups enabled us to reproduce the experimental pKa2 value. Furthermore, we demonstrated that specific conformations of acid/base pair of 5-HT is critical to predict accurately the experimental pKa’s of the flexible 5-HT molecule.  相似文献   

20.
Effect of additives (A) comprising main group elements (K, P) and transition metals (Ni, Cr, Mo) to VOx/SiO2 catalysts on their catalytic behavior in oxidative dehydrogenation (ODH) of propane are studied. The experimental findings are compared with results of quantum chemical calculations performed for a model O=V---O---V* cluster in which V* atom is replaced by an atom of A. The selectivities to propene found in experiments change on introducing the additives, decreasing in the sequence VK(64)>VCr(38)>VMo=VNi(32)>VP(28)V0(27). Quantum chemical calculations indicate the change in the electron density on the atoms of the active O=V---O groups for the clusters with the additives’ atoms. The extent of the electron transfer from A towards the active group (the increase in the centre basicity) follows the sequence of the decreasing selectivity to propene, confirming the role of acid–base properties in controlling the selectivity in ODH of propane. Results of calculations show also an exceptional behavior of the potassium (the most effective additive for the increase in the selectivity in experiments); in contrast to all other additivies, K transfers electrons towards both V and O atoms in the active group, and does not form a definite O---K bond, but is rather “adsorbed” on the cluster.  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