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.
Roy KI  Lucy CA 《Electrophoresis》2003,24(3):370-379
The mobilities of a series of aromatic ammonium ions, ranging in charge from +1 to + 3, were investigated by capillary electrophoresis using buffers consisting of 0-75% v/v methanol. This is an extension of our previous studies involving anion mobility in methanol-water media [1]. Absolute mobilities were determined by extrapolation of the effective mobilities to zero ionic strength according to the Pitts' equation. For all of the buffer compositions studied, the ionic strength effect increased with increasing cation charge, and varied as a function of solvent 1/eta epsilon (1/2) as predicted by the electrophoretic term within the Pitts' equation. In the presence of methanol, the ionic strength effects became more dramatic. The absolute mobilities of the cations were altered by the addition of methanol to the electrophoretic media. For example, at 75% MeOH, a migration order reversal was observed between the + 2 and + 3 ammonium ions. These solvent-induced selectivity changes are attributed to dielectric friction. As predicted by the Hubbard-Onsager dielectric friction model, dielectric friction increased with increasing methanol content and with increasing analyte charge. Further, the changes in cation mobility correlated to the changes in solvent relaxation time (tau), epsilon and eta. Although not predicted by the Hubbard-Onsager theory, the + 3 ammonium ion experienced more dielectric friction than the - 3 sulfonate and - 3 carboxylate investigated previously [1]. This apparent failure of the Hubbard-Onsager model results from its continuum nature, whereby ion-solvent interactions are not taken into account.  相似文献   

2.
The mobilities of three aromatic sulfonates, ranging in charge from -1 to -3, were investigated by capillary electrophoresis using buffers containing 0 to 75% ethanol or 2-propanol. Absolute mobilities were determined by extrapolation of the effective mobilities to zero ionic strength according to the Pitts' equation. For all buffers studied, ions of higher charge experienced larger ionic strength effects. The resulting ionic strength-induced selectivity alterations were more dramatic when organic solvents were present in the media. Furthermore, for different organic modifier types and contents, the magnitude of the ionic strength effect was governed to a large extent by the 1/(eta epsilon 1/2) dependence in the electrophoretic effect of the Pitts' equation. Addition of ethanol or 2-propanol to the electrophoretic media resulted in changes in the absolute mobilities of the ions. These solvent-induced mobility changes are attributed to dielectric friction. As predicted by the Hubbard-Onsager model, dielectric friction increased with increasing organic content and with increasing analyte charge. As a result, dramatic changes in the relative absolute mobilities were observed, such as a reversal in migration order between sulfonates of -1 and -3 charge in 75% 2-propanol. Within the alcohols, the Hubbard-Onsager model was successful at predicting the relative mobility trends upon changing solvent. However, the relative trends observed between acetonitrile-water and alcohol-water media were not consistent with the model. This may be explained by the continuum nature of the model, whereby the different ion-solvent interactions characteristic to each solvent class are not taken into account.  相似文献   

3.
When modeling analyte-additive interactions in capillary electrophoresis (CE), it is necessary to correct for all changes in the apparent electrophoretic mobility of an analyte that are not due to specific binding. Current models based on dynamic complexation have corrected for bulk viscosity changes in the background electrolyte (BGE) when additives are used, while assuming negligible changes in the dielectric constant and other physicochemical properties of the solution. In this report, a study of weak interactions between deoxyribonucleotides and hydroxypropyl-beta-cyclodextrin (HP-beta-CD) revealed significant nonideality in binding isotherms. Changes in the dielectric properties of the solution due to the addition of high concentrations of HP-beta-CD to the BGE was observed to alter the electrophoretic mobility of analytes. A relative dielectric correction factor was required to normalize analyte mobilities to a reference state of zero additive concentration. The use of both a relative dielectric factor and a viscosity correction factor was found to increase the accuracy of the model, reflected by a higher degree of correlation between predicted and measured analyte mobilities. This type of correction is particularly relevant when studying weak analyte binding interactions or when using high concentrations of additive in the BGE. This work is vital for accurate determination of weak binding constants and mobility values, as well as providing a deeper understanding of the fundamental parameters influencing a separation in CE.  相似文献   

4.
The free solution mobilities of the adenosine nucleotides 5'-adenosine triphosphate (ATP), 5'-adenosine diphosphate (ADP), 5'-adenosine monophosphate (AMP), and 3'-5'-cyclic AMP (cAMP) have been measured in diethylmalonate buffers containing a wide variety of monovalent cations. The mobilities of all nucleotides increase gradually with the increase in intrinsic conductivity of the cation in the BGE. However, at a given conductivity, the mobilities observed for ATP, ADP, and AMP in BGEs containing alkali metal ions and other cations are lower than these observed in BGEs containing tetraalkylammonium ions. Since the mobility of cAMP is independent of the cation in the BGE, the results suggest that the relatively low mobilities observed for ATP, ADP, and AMP in BGEs containing cations other than a tetraalkylammonium ion are due to cation binding, reducing the effective net charge of the nucleotide and thereby reducing the observed mobility. To measure the binding quantitatively, the mobilities of the nucleotides were measured as a function of ionic strength. The mobilities of ATP, ADP, and AMP decrease nonlinearly with the square root of ionic strength (I(1/2)) in BGEs containing an alkali metal ion or Tris(+). By contrast, the mobilities decrease linearly with I(1/2) in BGEs containing a nonbinding quaternary ammonium ion, as expected from Debye-Hückel-Onsager (DHO) theory. The mobility of cAMP, a nonbinding analyte, decreases linearly with I(1/2), regardless of the cation in the BGE. Hence, a nonlinear decrease of the mobility of an analyte with I(1/2) appears to be a hallmark of counterion binding. The curved mobility profiles observed for ATP, ADP, and AMP in BGEs containing an alkali metal ion or Tris(+) were analyzed by nonlinear curve fitting, using difference mobility profiles to correct for the effect of the physical properties of BGE on the observed mobilities. The calculated apparent dissociation constants range from 22 to 344 mM, depending on the particular cation-nucleotide pair. Similar values have been obtained by other investigators, using different methods. Interestingly, Tris(+) and Li(+) bind to the adenosine nucleotides with approximately equal affinities, suggesting that positively charged Tris(+) buffer ions can compete with alkali metal ions in Tris-buffered solutions.  相似文献   

5.
Li N  Duan J  Chen H  Chen G 《Talanta》2003,59(3):493-499
The apparent electrophoretic mobilities of procaine hydrochloride (μi) in a series of concentration of β-cyclodextrin were measured directly by capillary electrophoresis technology. A new mathematical treatment method is proposed, which based on the fact that the molar ratio of the inclusion complex was 1:1 established by spectrophotometry. Using the proposed method, the binding constant of the inclusion complex of procaine hydrochloride with β-cyclodextrin can be obtained easily. The determination result was in correspondence with those of the spectrophotometric and fluorescence methods.  相似文献   

6.
Underhill DW 《Talanta》1993,40(2):139-142
Some commercially available diffusive samplers use two layers of adsorbent placed in series. After sampling is completed, the time weighted average concentration of analyte is estimated from the weighted sum of the uptake of analyte on these two layers. It is known that such a division into layers can increase the permissible sampling time. Here the principles underlying this sampling procedure are analyzed through a fundamental application of the theory of diffusion. Using a trial and error procedure, the optimal division of adsorbent was calculated, and the increase in sampling time that such a division can give was confirmed theoretically. Also, should the uptake in the backup layer exceed a predetermined fraction of the total uptake, this will indicate misuse of the diffusive sampler.  相似文献   

7.
《Electrophoresis》2017,38(8):1188-1200
To optimize chiral separation conditions and to improve the knowledge of enantioseparation, it is important to know the binding constants K between analytes and cyclodextrins and the electrophoretic mobilities of the temporarily formed analyte‐cyclodextrin‐complexes. K values for complexes between eight phenethylamine enantiomers, namely ephedrine, pseudoephedrine, methylephedrine and norephedrine, and four different β‐cyclodextrin derivatives were determined by affinity capillary electrophoresis. The binding constants were calculated from the electrophoretic mobility values of the phenethylamine enantiomers at increasing concentrations of cyclodextrins in running buffer. Three different linear plotting methods (x ‐reciprocal, y ‐reciprocal, double reciprocal) and nonlinear regression were used for the determination of binding constants with β‐cyclodextrin, (2‐hydroxypropyl)‐β‐cyclodextrin, methyl‐β‐cyclodextrin and 6‐O‐α‐maltosyl‐β‐cyclodextrin. The cyclodextrin concentration in a 50 mM phosphate buffer pH 3.0 was varied from 0 to 12 mM. To investigate the influence of the binding constant values on the enantioseparation the observed electrophoretic selectivities were compared with the obtained K values and the calculated enantiomer‐cyclodextrin‐complex mobilities. The different electrophoretic mobilities of the temporarily formed complexes were crucial factors for the migration order and enantioseparation of ephedrine derivatives. To verify the apparent binding constants determined by capillary electrophoresis, a titration process using ephedrine enantiomers and β‐cyclodextrin was carried out. Furthermore, the isothermal titration calorimetry measurements gave information about the thermal properties of the complexes.  相似文献   

8.
The free solution mobilities of six single-stranded 16-nucleotide DNA oligomers with the same sequence, containing up to 11 neutral phosphoramidate internucleoside linkages, have been measured by capillary electrophoresis. The mobilities of the partially charged oligomers increase linearly with the logarithm of increasing charge density, as predicted by the Manning theory of electrophoresis (G. S. Manning, J. Phys. Chem. 1981, 85, 1506-1515). For comparison, the mobilities of eight fully charged single-stranded oligomers containing similar numbers of charged phosphate residues have also been measured. The mobilities of the variable length, fully charged oligomers increase more rapidly with the increasing number of charged phosphate residues than the mobilities of the constant size, partially charged phosphoramidate derivatives, because of the larger diffusion coefficients of the modified oligomers.  相似文献   

9.
《Chemical physics letters》1980,71(1):158-163
The rotational diffusion tensors of several rigid molecules are determined from the rotational friction tensor. The molecular friction tensor is modeled as a sum over the translational Stokes law friction coefficients of the constituent atoms of the molecule. Agreement of the theory with experimental rotational diffusion constants and onentational correlation times is surprisingly good.  相似文献   

10.
We study the rotational diffusion of a spherical colloid confined in a narrow channel between parallel plane hard walls. The walls damp translational diffusion much more than rotational diffusion so that there is expected to be little translation-rotation coupling. Using a recent calculation of the nonisotropic rotational mobilities arising from the hydrodynamic interactions with the walls, we set up the rotational Smoluchowski equation for either a particle with a permanent dipole moment or a polarizable particle with axisymmetric polarizabilities subject to an external electric field. Using the Smoluchowski equation dynamics we calculate the time-correlation functions of orientation that are measured in depolarized light scattering for the cases of no external field, external field normal to the walls, and external field parallel to the walls. The decay of correlations is shown to be given by a weighted sum of decaying exponentials and can be characterized by an initial and a mean characteristic decay time. The weights and decay rates of each component and the characteristic decay times are studied numerically for a range of field strengths. The nonisotropic rotational mobilities make these decay times highly sensitive to the distance of the particle from the confining walls. This position dependence can be used as a method of measuring the rotational mobilities or, conversely, the rate of decay of correlations can be used as a probe of particle position between the confining walls.  相似文献   

11.
Peak dispersion effects in nonaqueous capillary electrophoretic separations of aromatic anionic analytes were investigated in a propanolic background electrolyte solution. Poly(glycidylmethacrylate-co-N-vinylpyrrolidone) coating was applied to the capillary to suppress the electroosmotic flow and to improve the repeatability of the migration times. Electrical field strengths up to 2000 Vcm(-1) were applied in separations and the separation efficiencies were compared with theoretical values calculated on the basis of plate height theory. The contributions to the total plate height were calculated for injection plug length, diffusion, Joule heating, electromigration dispersion, analyte adsorption to the capillary wall, and detector slit aperture length. Analyte diffusion coefficients were measured by Taylor dispersion method, while distribution constants were measured chromatographically. Agreement between the calculated and empirical results was fairly good even though some approximations were required. In most cases the longitudinal diffusion contribution governed the total plate height, while the contribution of Joule heating was insignificant even at exceptionally high field strengths used. The relatively long detection slit aperture was found to influence the separation efficiency strongly, while the other dispersion sources that were investigated were of minor importance, except for adsorption in the case of one analyte. With all analytes, the dispersive effect of longitudinal diffusion was reduced as the field strength was increased, leading to enhanced migration velocities and faster separations.  相似文献   

12.
Catecholamines were analysed in aqueous and alcoholic non-aqueous solutions by capillary electrophoresis and capillary electrophoresis-mass spectrometry using sheathless nanospray coupling. Decreases in the electrophoretic mobilities of the catecholamines and in the electroosmotic mobilities were observed from water to 1-propanol. Separations were more efficient in all non-aqueous media than in water. The diffusion coefficients of the catecholamines in the different media were determined. The solvent had little effect on the sensitivity of the UV or MS detection. Both methods were successfully applied to the analysis of urine samples.  相似文献   

13.
Porras SP  Kenndler E 《Electrophoresis》2005,26(17):3279-3291
N,N-Dimethylformamide (DMF) is a dipolar protophilic solvent with physicochemical properties that makes it suitable as solvent for capillary electrophoresis (CE). It is prerequisite for the proper application of CE to adjust and to change the pH of the background electrolyte (BGE) in a defined manner. This was done in the present work using benzoic acid-benzoate by selecting different concentration ratios of acid and salt, and calculating the theoretical pH from the activity-corrected Henderson-Hasselbalch equation. The mobilities of the analytes (chloro- and nitro-substituted phenolates) were found to follow reasonably well the typical sigmoid mobility versus pH curve as predicted by theory. The actual mobilities and pK(a) values (at 25 degrees C) of the analytes were derived from these curves. pK(a) values were in the range of 11.1-11.7, being thus 3-4.4 units higher than in water. This pK(a) shift is caused by the destabilization of the analyte anion and the better stability (solubility) of the molecular analyte acid in DMF, which overcome the higher basicity of DMF compared to water. Absolute mobilities were calculated from the actual mobilities; they were between 32x10(-9) and 42x10(-9) m(2)/Vxs. Slight deviations of the measured mobilities from the theoretical mobility versus pH curve were discussed on the bases of ion pairing and heteroconjugation and homoconjugation of either buffer components or buffer components and analytes. Heteroconjugation was used as a mechanism for the electrically driven separation of neutral analyte molecules in a BGE where salicylate acted as complex forming ion. Rough estimation of the complexation constants for the phenolic analytes gave values in the range of 100-200 L/mol. Addition of water to the solvent decreased the effect of heteroconjugation, but it was still present up to the surprisingly high concentration of 20% water. Electrophoretically relevant parameters like ionic mobilities and pK(a) values, and conjugation and ion pairing are dependent on the water content of the solvent. The water uptake of DMF was measured when exposed to humidity of ambient air. The resulted behavior of the water uptake was found rather similar to that for acetonitrile and methanol.  相似文献   

14.
The electrophoretic mobilities of three beta-blocker drugs, practolol, timolol and propranolol, have been measured in electrolyte systems with mixed binary and ternary water-methanol-ethanol solvents with acetic acid/sodium acetate as buffer using capillary electrophoresis. The highest mobilities for the analytes studied have been observed in pure aqueous, the lowest values in ethanolic buffers. The measured electrophoretic mobilities have been used to evaluate the accuracy of a mathematical model based on a mixture response surface method that expresses the mobility as a function of the solvent composition. Mean percentage error (MPE) has been computed considering experimental and calculated mobilities as an accuracy criterion. The obtained MPE for practolol, timolol and propranolol in the binary mixtures are between 0.9 and 2.6%, in the ternary water-methanol-ethanol solvent system the MPE was about 2.7%. The MPE values resulting from the proposed equation lie within the experimental relative standard deviation values and can be considered as an acceptable error.  相似文献   

15.
The relatively new technique of reversed-flow gas chromatography (RFGC) is used to determine the diffusion coefficients of pure gases into gas mixtures (D(mix)(exp)). The pure gases are CO and CO(2), and the mixtures consist of H(2) and He in various volume percentage compositions. A linear regression analysis of D(mix)(exp) of CO and CO(2) in various mixtures of H(2) and He against the percentage composition (X(H2) or X(He)) of the mixtures at different temperatures results in an empirical equation relating D(mix)(exp) to the corresponding theoretical values of the diffusion coefficients of CO and CO(2) in the pure gases H(2) and He, as they are calculated from the Fuller-Schettler-Giddings equation. The empirical equation shows that the diffusion coefficient of an analyte gas in a gas mixture is the partial sum of its diffusion coefficients in the component gases, therefore making possible the determination of the mole fractions of the components of the mixture. The found percentage volume compositions are very close to those determined independently by routine gas chromatography, indicating that the proposed RFGC methodology could be successfully applied to the accurate determination of the volume composition of binary gas mixtures.  相似文献   

16.
Two selectors based on modified cyclodextrins containing a metal binding site and a dansyl fluorophore-6-deoxy-6-N-(N(alpha)-[(5-dimethylamino-1-naphthalenesulfonyl)aminoethyl]phenylalanylamino-beta-cyclodextrin-containing D-Phe (3) and L-Phe (4) moieties were synthesized. The conformations of the two selectors were studied by circular dichroism, two-dimensional NMR spectroscopy and time-resolved fluorescence spectroscopy. Cyclodextrin 4 was found to have a predominant conformation in which the dansyl group is self-included in the cyclodextrin cavity, while 3 showed a larger proportion of the conformation with the dansyl group outside the cavity. As a consequence, the two cyclodextrins were found to bind copper(II) with different affinities, as revealed by fluorescence quenching in competitive binding measurements. Addition of D- or L-amino acids induced increases in fluorescence intensity, which were dependent on the amino acid used and in some cases on its absolute configuration. The cyclodextrin 4 was found to be more enantioselective than 3, suggesting that the self-inclusion in the cyclodextrin cavity strongly increases the chiral discrimination ability of the copper(II) complex. Accordingly, a linear fluorescent ligand N(alpha)-[(5-dimethylamino-1-naphthalenesulfonyl)aminoethyl]-N(1)-propyl-phenylalaninamide, which has the same binding site and absolute configuration as 4, showed very low chiral discrimination ability. The enantioselectivity in fluorescence response was found to be due to the formation of diastereomeric ternary complexes, which were detected by ESI-MS and by circular dichroism. Time-resolved fluorescence studies showed that the fluorescence of the dansyl group was completely quenched in the ternary complexes formed, and that the residual fluorescence was due to uncomplexed ligand.  相似文献   

17.
Distortion of the starting zone upon its electrophoretic migration toward the detection window gives rise to both symmetrical zones caused by diffusion, sedimentation in the horizontal section of the capillary and the curvature of the capillary, and asymmetrical zones having their origin in Joule heating, sedimentation in the vertical section of the capillary, pH and conductivity differences between the sample zone and the surrounding buffer, solute adsorption onto the capillary wall, and association-dissociation of complexes between the analyte and a buffer constituent or between analytes. Interestingly and importantly a theoretical study shows that moderate pH and conductivity differences as well as adsorption and all of the above interactions when they are characterized by a fast on/off kinetics do not increase the zone broadening (or only slightly), because the sharpening of one boundary of the zone is about the same as the broadening of the other boundary. In addition the peak symmetry caused by a conductivity difference is in most experiments counteracted by a pH difference. The experimentally determined plate numbers in the absence of electroosmosis exceeded one million per meter in some experiments (Part II). These plate numbers are among the highest reported [Z. Zhao, A. Malik, M.L. Lee, Anal. Chem. 65 (1993) 2747; M. Gilges, K. Kleemiss, G. Schomburg, Anal. Chem. 66 (1994) 2038; H. Wan, M. Ohman, L.G. Blomberg, J. Chromatogr. A 924 (2001) 591 (plate numbers determined in the presence of electroosmosis may be higher, although the width of the zone in the capillary may be larger) [p. 680 in S. Hjertén, Electrophoresis 11 (1990) 665]). Capillary free zone electrophoresis is perhaps the only separation method, which, under optimum conditions, gives a plate number not far from the theoretical limit. A prerequisite for this high performance is that the polyacrylamide-coated capillary is washed with 2 M HCl between the runs and stored in water over night (Part II). The difference between the experimentally determined total variance and the sum of the calculated variances originating from the width of the starting zone, longitudinal diffusion, Joule heating, sedimentation in the vertical section of the capillary, curvature of the capillary (i.e., the sum of all other variances) was in our most successful experiments about 28% of the variance of diffusion. The zone broadening, 2sigma, caused by diffusion was estimated at 0.77 mm. The total zone width (2sigma) calculated from the experimentally determined plate number was as small as 1 mm when the migration distance was 40 cm. Accordingly, the only efficient way to reduce drastically the total zone width is to decrease the analysis time and, thereby, the diffusional broadening. An important finding was that the variance originating from the loops of the capillary is not always negligible in high-performance runs. Therefore, one should employ straight capillaries and avoid CE apparatus with cartridges that require a strong curvature of the capillary, common in most commercial instruments. Mathematical formulas have been derived for the sedimentation of the solute zone, the enrichment factor, and the migration time in experiments where the solute is dissolved in a dilute running buffer. This zone sharpening method gave very narrow starting zones (0.04-0.4 mm). However, upon high dilution of the buffer the enrichment becomes so strong that part of the sample zone probably sediments out of the capillary; the almost inevitable change in pH may decrease the mobility of the proteins and, thus, cause the enrichment factor to become still lower than expected. Diffusion of the protein in the very narrow starting zone (located close to the tip of the capillary) and sometimes the thermal expansion of the buffer in the capillary contributes to additional loss of protein in the enrichment step. In some buffers, the interaction between the protein and the buffer constituents is so slow that the peaks become broad. Therefore, different types of buffers should be tested when high resolution is required. The relation sigma2 (the variance of the interaction between a protein and the buffer constituents) = constant x u (the mobility) seems to be valid for all proteins in the applied sample, at least when they have similar molecular masses. To facilitate the understanding of the progress of a free zone electrophoresis experiment, we have discussed in simple terms how the concentrations of the background electrolytes become rearranged during a run and why the difference between the mobilities of the proteins and the mobilities of the background electrolyte determines whether a peak exhibits fronting or tailing. A theoretical analysis of zone broadening in capillary zone electrophoresis, chromatography, and electrochromatography indicates that electrochromatography in homogeneous gels might be the only chromatographic technique which can compete in performance with free electrophoresis. Using an equation, valid not only for electrophoresis, but also for chromatography and centrifugation, the mobility of a concentration boundary has been calculated for the first time and was, as expected, low. Equations based on the Kohlrausch regulating function do not permit such calculations. Another regulating function (the H function) and some of its characteristics are briefly discussed. The theoretical discussions in this paper and the experimental studies in Part II show that high-performance electrophoresis deserves its prefix when the runs are designed to give minimum zone broadening. Some guidelines are given to facilitate this optimization. The plate numbers are so high that the resolution cannot be increased by more than 30% even if they approach the theoretically maximum values.  相似文献   

18.
The stability of the molecular host-guest inclusion complexes of beta-cyclodextrin with benzoate and four different hydroxybenzoates is investigated. For the measurement of the binding constants an experimental method is devised that is based on affinity capillary electrophoresis (ACE) with indirect UV absorbance detection. We derive an explicit equation for effective mobilities in ACE experiments without violation of rigorous mass balance. This equation is employed in the nonlinear least-squares analyses of the experimental data yielding binding constants of 48+/-2 M(-1) for benzoate, 299+/-38 M(-1) for 2-hydroxybenzoate, 37+/-1 M(-1) for 3-hydroxybenzoate, 228+/-9 M(-1) for 4-hydroxybenzoate, and 895+/-110 M(-1) in the case of 2,4-dihydroxybenzoate.  相似文献   

19.
Maichel B  Gas B  Kenndler E 《Electrophoresis》2000,21(8):1505-1512
Apparent diffusion coefficients, Dapp,i, were determined in solutions with a polycationic additive -- poly(diallyidimethylammonium) -- acting as a pseudostationary phase for electrokinetic chromatography. They were determined for six small neutral analytes at five concentrations of the polymeric additive (between 0 and 4% w/w) by a stopped migration method. The apparent diffusion coefficients decrease with increasing polymer concentration only within 40% maximum, an effect that cannot be associated with the macroscopic viscosity of the polymer solution (which increases by a factor of 10). The change of the apparent diffusion coefficients is related to the interaction of the neutral analyte molecules with the polyelectrolyte chain. Applying the model of analyte partitioning between "free" solution and polymer, capacity factors and partition constants were derived from the slope of the 1/Dapp,i vs. polymer concentration curves. Partition constants determined by this method (ranging between 40 and 170) agree with those obtained by electrokinetic chromatography.  相似文献   

20.
Hsiao JY  Wu SH  Ding WH 《Talanta》2006,68(4):1252-1258
Application of capillary electrophoresis (CE) to simultaneously determine the apparent binding constants and thermodynamic parameters for six positional and structural naphthalenesulfonate derivatives with β-cyclodextrin (β-CD) is presented. The change in electrophoresis mobilities was used to assess the binding constants by non-linear regression and three different linear plots methods (named double reciprocal, x-reciprocal and y-reciprocal). The substituent group(s) attached to the naphthalene ring considerably affected the inclusion behaviors of these naphthalenesulfonate derivatives. The binding constant varies over almost one order of magnitude and a highly selective sequence is obtained between these guest model compounds. Naphthalenesulfonates with the substituent(s) at the 2-position(s) displayed stronger interaction with β-CD, and gave well compatible results by these four plot methods. While at least one substituent was substituted into the 1-position of naphthalene showed the weak interaction or no interaction with β-CD. Comparison to three linear regression methods, the non-linear regression method proves to be the most suitable for these determinations. Additionally, apparent binding constants for each structural isomer with β-CD at several temperature, and thermodynamic parameters for binding were also calculated and discussed.  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