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 546 毫秒
1.
The determination of the diuretics hydrochlorothiazide, bendroflumethiazide and furosemide by both conventional and thermal lens spectrophotometry (TLS, 100 mW of pump power at 514.5 nm) following previous hydrolysis, diazotization and coupling with N-(naphthyl)ethylenedine (NED) in a sodium dodecyl sulphate (SDS) micellar medium of pH approximately 1 was studied. p-Aminobenzoic acid (PABA) was used as a model compound to optimize the derivatization procedures. 3-Substituted indoles, such as 5-hydroxyindole-3-acetic acid and tryptophan, gave N-nitroso derivatives which interfered with the determination of the diuretics in urine. The derivatized diuretics in urine were separated using HPLC with a Spherisorb ODS-2 C(18) column, and a 0.1M SDS mobile phase containing 5% n-propanol and 0.001M sodium dihydrogen phosphate (pH 3). The diuretics gave limits of detection (LODs) of ca. 5 x 10(-9)M for the TLS procedure. The LODs were 20-50-fold higher for the corresponding spectrophotometric procedure.  相似文献   

2.
The performance of dopant-assisted atmospheric pressure photoionization (DA-APPI) and electrospray ionization (ESI) for the coupling of micellar electrokinetic chromatography (MEKC) with ion trap mass spectrometry (ITMS) was compared using a set of test drugs comprising basic amines, steroids, esters, phenones and a quaternary ammonium compound. The influence of the surfactant sodium dodecyl sulfate (SDS) on analyte signals was studied by infusion of sample through the CE capillary into the respective ion sources. It was found that background electrolytes (BGEs) containing 20-50mM SDS in 10mM sodium phosphate (pH 7.5) caused major ionization suppression for both polar and apolar compounds in ESI-MS, whereas APPI-MS signal intensities remained largely unaffected. ESI gave rise to the formation of SDS clusters, which occasionally may cause space-charge effects in the ion trap. Furthermore, extensive sodium-adduct formation was observed for medium polar compounds with ESI-MS, whereas these compounds were detected as their protonated molecules with APPI-MS. Using the BGE containing 20mM SDS, MEKC-ESI-MS still provides slightly lower limits of detection (LODs) (2.6-3.1muM) than MEKC-APPI-MS (4.3-6.4muM) for basic amines. For less polar compounds, highest S/Ns were obtained with APPI-MS detection (LODs, 4.5-71muM). For BGEs containing 50mM SDS, the limits of detection for MEKC-APPI-MS were more favorable (factor 1.5-12) than MEKC-ESI-MS for nearly all tested drugs. Spray shield contamination by SDS was lower in DA-APPI-MS than in ESI-MS. It is concluded that DA-APPI shows the most favorable characteristics for MEKC-MS, especially when compounds of low polarity have to be analyzed.  相似文献   

3.
The binding of the photosensitizing fluoroquinolone (FQ) antibiotic norfloxacin (NX) to sodium dodecyl sulfate (SDS) micelles and the photoreactivity of the NX/SDS complex under physiological pH conditions are investigated by means of absorption and emission spectroscopy, steady-state and laser flash photolysis. It is shown that the photolabile zwitterionic form of NX, which is dominant at physiological pH, is not the most abundant species in the presence of SDS micelles. This medium exhibits a high preference for the cationic form of the drug, which is selectively and successfully entrapped within the micellar cage (K(ass) = 6 x 10(4) M(-1) +/- 3000), becoming the largely dominant species at neutral pH. The effect of this trapping is drastically reflected on both efficiency and nature of the drug photodecomposition. It is observed that the photostability of NX incorporated in the micellar pseudophase increases of more than one order of magnitude if compared to that of the "free" drug. Furthermore, the radical photodecomposition mechanism occurring in phosphate buffered solution is suppressed by the micellar medium and the low photodegradation observed seems to take place preferentially through an ionic pathway. Hopefully, the results presented herein may contribute to a better understanding of the bio-distribution of NX in biological systems and provide helpful and stimulating information in order to get the control of FQ photoreactivity under physiological pH conditions.  相似文献   

4.
CE was coupled to inductively coupled plasma MS (ICP-MS) and ESI-MS to identify and quantify the arsenic species arsenobetaine (AsB), arsenite (As(III)), arsenate (As(V)), and dimethylarsinic acid (DMA). A GC-flame ionization detector (FID)-based German standard method and ICP-MS were used for validation of the data obtained for arsenobetaine and total arsenic, respectively. LODs obtained with the CE-ESI-TOF-MS method were 1.0x10(-7) M for AsB, 5.0x10(-7) M for DMA, and 1.0x10(-6) M for As(III) and As(V). For the CE-ICP-MS method, LODs were 8.5x10(-8) M for AsB, 9.5x10(-8) M for DMA, 9.3x10(-8) M for As(III), and 6.2x10(-8) M for As(V). While CE-ICP-MS provided high sensitivity and better reproducibility for quantitative measurements, CE-ESI-MS with a TOF mass analyzer proved to be valuable for species identification. With this setup, fish samples were prepared and analyzed and the obtained data were successfully validated with the independent methods.  相似文献   

5.
Huang X  Zhang W  Han S  Yin Y  Xu G  Wang X 《Talanta》1997,45(1):127-135
Different micellar media had different effects on the absorption spectra of the complexes of bromopyrogallol red with Sb(III) and Sb(V). The mixed micellar medium composed of 0.7 ml of 0.2% sodium dodecylsulfate (SDS) and 0.3 ml of 2% nonylphenoxypolyethoxyethanol (OP) at 80 degrees C could be used for the sensitive determination of Sb(III) in Sb(III)/Sb(V) binary mixtures. Under the optimal conditions, Beer's Law was obeyed over the range 0.1-2.3 mug ml(-1) Sb(III) with molar absorptivity at 538 nm being 4.8 x 10(4) l mol(-1) cm(-1) and detection limit 0.04 mug ml(-1). For 10 mug Sb(III), more than 100 mug Sb(V) could be tolerated (error < 3%) in the presence of SDS/OP micellar medium as compared with 0.1 mug Sb(V) in the absence of SDS/OP micellar medium. In addition, the sensitivity of Sb(III) in the micellar medium was much higher than that in pure water medium. As compared with conventional extraction spectrometry, the proposed method produced a reproducible result. It did not need the conversion of Sb(III) to Sb(V) and a time-consuming extraction process. A detailed discussion on the selection of surfactants, the effect of temperature, and the role played by the mixed surfactants were also made.  相似文献   

6.
Wu YW  Jiang YY  Liu JF  Xiong K 《Electrophoresis》2008,29(4):819-826
A method has been developed for the separation and determination of three hydrophobic benzophenones: 2,4-dihydroxybenzophenone (BP-1), 2,2'4,4'-tetrahydroxybenzophenone (BP-2), and 2-hydroxy-4-methoxybenzophenone (BP-3) in sunscreen by micellar electrokinetic capillary chromatography (MEKC) combined with cloud point extraction (CPE). The analytes were extracted at pH 5.0 by micelles of the nonionic surfactant polyoxyethylene-7.5-octylphenyl ether (Triton X-114). A 150 microL aliquot from the extracted surfactant-rich phase was diluted up to 500 microL with ethanol to reduce its viscosity before separation by MEKC. A background electrolyte of 25 mmol/L sodium tetraborate containing 30 mmol/L sodium dodecyl sulfate at pH 9.25 was used as the separation medium to avoid the adsorption of hydrophobic substances and Triton X-114 onto the inner surface of the separation capillary, ensuring the separation efficiency and reproducibility. Detection is performed at 290 nm. Under the optimized conditions, an enrichment factor of 20 was obtained and the determination limits of BP-1, BP-2, and BP-3 were found to be 3.90 x 10(-) (7), 3.83 x 10(-7), and 6.42 x 10(-8) mol/L, respectively. In comparison with the earlier reported methods, the LODs of this method are superior to the other methods. The presented procedure was successfully applied to the determination of BP-1, BP-2, and BP-3 in sunscreen with satisfactory results.  相似文献   

7.
A sensitive flow-injection (FI) procedure with spectrophotometric detection in a micellar medium is proposed for the determination of novalgin. The method is based on the instantaneous formation of a red-orange product (lambda(max) = 510 nm) after the reaction between novalgin and p-dimethylaminocinnamaldehyde (p-DAC) in a dilute acid medium. The sensitivity of this reaction was increased by a factor of 5.6 in the presence of sodium dodecyl sulfate (SDS). Experimental design methodologies were used to optimize the chemical and FI variables. The calibration curve was linear in the range of 1.45 x 10(-6) to 2.90 x 10(-5) mol L(-1) with an excellent correlation coefficient (r = 0.9999). The detection limit was 1.31 x 10(-7) mol L(-1) (n = 20, RSD = 2.0%). No interferences were observed from the common excipients. The results obtained by the proposed method were favorably compared with those given by the iodometric reference method at 95% confidence level.  相似文献   

8.
An analytical procedure has been developed for the separation of dioxouranium(VI), iron(III), copper(II), nickel(II), cobalt(II), cobalt(III), palladium(II), and thorium(IV) by MEKC using N,N'-ethylene bis(salicylaldimine) (H(2)SA(2)en) as a complexing reagent with total runtime <4.5 min. SDS was used as micellar medium at pH 8 with sodium tetraborate buffer (0.1 M). An uncoated fused-silica capillary with an effective length of 50 cm x 75 microm id was used with an applied voltage of 30 kV with photodiode array detection at 231 nm. Linear calibrations were obtained within 0.111-1000 microg/mL of each element with LODs within 37-325 ng/mL. The developed method was tested for analysis of uranium ore samples indicating its presence within 103-1789 microg/g with RSD within 0.79-1.87%. Likewise copper, nickel, and iron in their combined matrix were also simultaneously determined with RSD 0.4-1.6% (n = 6).  相似文献   

9.
We investigated the binding of sodium dodecyl sulfate (SDS) to various linear and star polymers of the nonionic methoxyhexa(ethylene glycol) methacrylate (PMHEGMA) and the ionic 2-(dimethylamino)ethyl methacrylate (PDMAEMA), the latter being a polycation at low pH. The dodecyl sulfate ion selective electrode (EMF), isothermal titration calorimetry (ITC), and surface tension (ST) were applied to gain detailed information about interactions. In all cases there is evidence of significant binding of SDS over an extensive SDS concentration range spanning from ca. 10(-6) to 0.1 mol dm(-3). At pH 3, the polymer PDMAEMA is a strong polycation and here the binding is dominated by electrostatic 1:1 charge neutralization with the anionic surfactant. At their natural pH of 8.6, PMHEGMA and PDMAEMA polymers are essentially nonionic and bind SDS in the form of polymer-bound aggregates in the concentration range of ca. 1 x 10(-3) to 3 x 10(-2) mol dm(-3). All the polymers also bind SDS to a lesser extent at concentrations below 1 x 10(-3) mol dm(-3) reaching as low as 10(-7) mol dm(-3). This low concentration binding process involves the polymer and nonassociated SDS monomers. As far as we are aware, this is the first example that such a low concentration noncooperative binding process could be observed in SDS/neutral polymer systems by EMF and ST. We also showed that the nonionic surfactant hexa(ethylene glycol) mono-n-dodecyl ether (C12EO6) and the cationic cetyltrimethylammonium bromide (C16TAB) interact with star PDMAEMA. We believe that the interaction of C12EO6 and CTAB is of similar noncooperative type as the first SDS binding process in the range from ca. 10(-5) to 0.3 x 10(-3) mol dm(-3). At the high concentration binding limit Csat of SDS, the above polymers become fully saturated with bound SDS micelles. We applied small angle neutron scattering (SANS) to determine the structure and aggregation numbers of the star polymer/bound SDS micelles and calculated the stoichiometry of such supramolecular complexes. The SANS data on PDMAEMA star polymers in the presence of C12EO6 showed only a limited monomer binding in contrast to linear PDMAEMA, which showed monomer C12EO6 binding at low concentrations but micellar aggregates at 6 x 10(-3) mol dm(-3).  相似文献   

10.
Nitrobenzenes, such as the antibiotic chloramphenicol, the vasodilator nicardipine, and the herbicides dinitramin, dinobuton, fenitrothion, methylparathion, oxyfluorfen, parathion, pendimethalin, quintozene, and trifluralin, were determined by using a spectrophotometric method in the visible region (540 nm). The method was based on the reduction of the nitrobenzenes to arylamines with tin(II) chloride, diazotisation of the arylamines and coupling of the diazonium ions with the Bratton-Marshall reagent. The two latter reactions were performed in a micellar medium of sodium dodecyl sulphate. The linear calibration range was 2x10(-6) to 7x10(-5) M (r>0.999), with limits of detection in the 10(-7) M level, which is 2-6 fold lower with respect to the corresponding spectrophotometric procedure in non-micellar medium. The procedure was applied to the analysis of the compounds in commercial preparations (pharmaceuticals and herbicide formulations) and in water samples, with good recoveries.  相似文献   

11.
The use of a micellar solution as a means for improving the colorimetric determination of arylamines by diazotization and coupling is studied. Sodium dodecylsulphate (SDS), Triton X-100 (TX-100) and N-cetylpyridinium chloride (NCPC), together with the diazotizable substances aniline and sulphanilic acid and the coupling agents 1-naphthylamine and N-(1-naphthyl)ethylenediamine are used as model compounds. The protonation constants of the dyes follow the order SDS>H2O>TX-100>NCPC. In an SDS micellar solution the coupling rate increases greatly, allowing the rapid formation of the protonated dyes in an acetate buffer. Other advantages are the possibility of analysing very non-polar samples and of using a wider range of diazotizable substances and substrates. A procedure for the determination of 10–100 μg of aniline in an SDS micellar solution is proposed. The procedure is applied to the determination of sulphonamides in pharmaceutical preparations.  相似文献   

12.
Abstract— The deactivation rate of excited pyrene by indole strongly depends on the polarity of the media. In micellar systems (Triton X-100, cetyltrimcthylammonium chloride (CTAC) and sodium dodecylsulfate (SDS) the deactivation efficiency is enhanced due to the high local concentration of indole in the micellar pseudophase. Quantitative interpretation of the data in CTAC and SDS micelles requires to take into account indole exchange between the micelles and the aqueous phase. In SDS micelles, where due to their smaller size the exchange process is more relevant, the exit and entrance rates are (3.0 ± 0.6) x 106 and (1.2 ± 0.3) x 1010 M −1s−1 respectively. Intramicellar bimolecular quenching constants are (1.1 ± 0.2) x 108 M−1 s−1 (1.4 ± 0.2) x 108 M −1 s−1 and (1.5 ± 0.2) x 108 M −1 s−1 in Triton X-100, SDS and CTAC respectively. These rates are similar to those measured in ethanol rich ethanol-water homogeneous solutions. This is in agreement with the average polarity sensed by both pyrene and indole in the micellar pseudophases.  相似文献   

13.
The synthetic flavylium salt 4-carboxy-7-hydroxy-4'-methoxyflavylium chloride (CHMF) exhibits two acid-base equilibria in the range of pH 1-8 in both aqueous and micellar sodium dodecyl sulfate (SDS) solutions. The values of pK(a1) and pK(a2) for the cation-zwitterion (AH(2)(+) <--> Z + H(+)) and the zwitterion-base (Z <--> A(-) + H(+)) equilibria increase from 0.73 and 4.84 in water to 2.77 and 5.64 in SDS micelles, respectively. The kinetic study of the Z <--> A(-) + H(+) ground-state reactions in SDS points to the diffusion-controlled protonation of A(-) in the aqueous phase (k(p2w) = 4.2 x 10(10) M(-)(1) s(-)(1)) and in the micelle (k(p2m) = 2.3 x 10(11) M(-)(1) s(-)(1)). The deprotonation rate of Z did not significantly change upon going from water (k(d2) = 6.3 x 10(5) s(-)(1)) to SDS (k(d2) = 5.2 x 10(5) s(-)(1)), in contrast with the behavior of ordinary cationic flavylium salts, for which k(d2) strongly decreases in SDS micelles. These results suggest that deprotonation of the zwitterionic acid is not substantially perturbed by the micellar charge. Electronic excitation of the Z form of CHMF induces fast adiabatic deprotonation of the hydroxyl group of Z() (2.9 x 10(10) s(-)(1) in water and 8.4 x 10(9) s(-)(1) in 0.1 M SDS), followed by geminate recombination on the picosecond time scale. Interestingly, while recombination in water (k(rec) = 1.7 x 10(9) s(-)(1)) occurs preferentially at the carboxylate group, at the SDS micelle surface, recombination (k(rec) = 9.2 x 10(9) s(-)(1)) occurs at the hydroxyl group. The important conclusion is that proton mobility at the SDS micelle surface is substantially reduced with respect to the mobility in water, which implies that geminate recombination should be a general phenomenon in SDS micelles.  相似文献   

14.
The separation of fourteen active ingredients used in a cold medicine was investigated by micellar electrokinetic chromatography (EKC) employing bile salts. Basic drugs were also successfully separated by micellar EKC using bile salts with high theoretical plate numbers (2.0 x 10(5)-3.5 x 10(5)) within a relatively short time (ca. 20 min). The separation of these solutes by micellar EKC was not successful using sodium dodecyl sulphate. The effects of micellar concentration, pH and organic modifier content on migration times and selectivity were investigated. This technique was also applied to the determination of several active ingredients combined in commercial preparations by an internal standard method.  相似文献   

15.
An electroanalytical method for the determination of the herbicide desmetryne at nanomolar levels in dispersed media, based on adsorptive stripping voltammetry, is reported. The adsorption of desmetryne at the hanging mercury drop electrode was checked both in micellar solutions, where the anionic surfactant sodium pentanesulphonate was chosen as the most suitable surfactant agent, and in oil-in-water emulsions prepared with ethyl acetate as the organic solvent. In a micellar medium formed with 0.02% sodium pentanesulphonate and with 0.1 mol l(-1) Britton-Robinson buffer (pH 1.5), the herbicide could be determined over the 1.0 x 10(-8)-4.0 x 10(-7) mol l(-1) concentration range, when an accumulation potential of -0.70 V was applied for 50 s. On the other hand, in an oil-in-water emulsion formed with 2% ethyl acetate and 0.04% sodium pentanesulphonate as emulsifying agent in 0.1 mol l(-1) HClO(4), desmetryne could be determined over the 2.0 x 10(-9)-1.0 x 10(-7) mol l(-1) concentration range. The limits of detection were 2.4 x 10(-9) and 4.2 x 10(-10) mol l(-1) in micellar and emulsified media, respectively, with R.S.D.s (n=10) 3.6 and 3.7%. The degree of interference from some other s-triazines on the desmetryne differential pulse response was also evaluated. Finally, the method developed in emulsified medium was applied to the determination of desmetryne in spiked apple juice.  相似文献   

16.
A novel method is proposed for the determination of trace amount of oxalic acid in the present article. In 1.0M hydrochloric acid medium, oxalic acid can react with the zirconium(IV) in Zr(IV)-(DBS-arsenazo) complex and replaces the DBS-arsenazo to produce a hyperchromic effect at 520 nm. The hyperchromic degree is proportional to the concentration of the oxalic acid added over a defined range. Based on this property, a new method for the spectrophotometric determination of trace oxalic acid was developed. Beer's law is held over the concentration range of 9.0 x 10(-6) to 5.0 x 10(-4)M for oxalic acid with a correlation coefficient of 0.9995. The apparent molar absorptivity of the method is epsilon520 nm = 1.16 x 10(3)L mol(-1)cm(-1) and the detection limit for oxalic acid is 0.815 microg/mL. The developed method was directly applied to the determination of oxalic acid in tomato samples with satisfactory results.  相似文献   

17.
Safavi A  Karimi MA 《Talanta》2002,58(4):785-792
A rapid and sensitive flow injection chemiluminescence (CL) method is described for the determination of hydrazine based on the CL generated during its reaction with either sodium dichloroisocyanurate (SDCC) or trichloroisocyanuric acid (TCCA) in alkaline medium. The emission intensity is greatly enhanced if dichlorofluorescein (DCF) as sensitizer is present in the reaction medium. The presence of citrate prevents the precipitation of some cations in the reaction medium and also causes an enhancement in emission intensity. The effect of analytical and flow injection variables on these CL systems and determination of hydrazine are discussed. The optimum parameters for the determination of hydrazine were studied and were found to be the following: SDCC and TCCA both 1x10(-3) M; NaOH, 2x10(-1) M; DCF, 5x10(-6) M; citrate, 1x10(-3) M and flow rate, 3.8 ml min(-1). The optimized method yielded 3sigma detection limits of 2x10(-7) and 3x10(-7) M for hydrazine with SDCC and TCCA oxidants, respectively. The method is simple, fast, sensitive, and precise and was applied to the determination of hydrazine in water samples.  相似文献   

18.
Su HL  Feng LI  Jen HP  Hsieh YZ 《Electrophoresis》2008,29(20):4270-4276
We have employed a high-sensitivity on-line preconcentration method, cation-selective exhaustive injection (CSEI) and sweeping MEKC, for the analysis of cocaine, benzoylecgonine, norcocaine, and cocaethylene. We monitored the effects of several of the CSEI-sweeping-MEKC parameters - including the pH, the concentrations of SDS and organic modifier, the injection length of the high-conductivity buffer, and the injection time of the sample - to optimize the separation process. The optimal BGE was 100 mM phosphoric acid (pH 1.8) containing 75 mM SDS with 10% 2-propanol and 10% tetrahydrofuran as the organic modifier. In addition, electrokinetic injection of the sample at 15 kV for 900 s provided both high separation efficiency and enhanced sweeping sensitivity. The sensitivity enhancements for cocaine, norcocaine, and cocaethylene ranged from 2.06 x 10(4) to 3.96 x 10(4); for benzoylecgonine it was 1.75 x 10(3); the coefficients of determination exceeded 0.9958. The LODs, based on an S/N ratio of 3:1, of sweeping-MEKC ranged from 33.5 to 52.8 ng/mL; in contrast, when using CSEI-sweeping-MEKC the sensitivity increased to range from 29.7 to 236 pg/mL. Under the optimal conditions, we analyzed cocaine in a human urine sample prepared using off-line SPE to minimize the influence of the matrix. The recovery of the SPE efficiency was satisfactory (ca. 74.9-87.6%). Our experimental results suggest that, under the optimal conditions, the CSEI-sweeping-MEKC method can be used to determine cocaine and its metabolites with high sensitivity in human urine.  相似文献   

19.
植物生长激素的毛细管胶束电动色谱法分离   总被引:5,自引:0,他引:5  
袁敏  张铭光  康经武  李菊白 《色谱》1997,15(6):482-485
以高效毛细管胶束电动色谱法对赤霉素(GA)、脱落酸(ABA)、吲哚丁酸(IBA)、吲哚乙酸(IAA)、萘乙酸(NAA)等植物生长激素的分离和测定进行了研究。考察了各种操作参数及有机添加剂对分离的影响,得到良好的分离结果。对各组分进行了定量测定研究,ABA、GA、IBA、IAA及NAA的最低检测浓度依次为5.0,3.0,0.58,0.15,0.14mg/L。  相似文献   

20.
Micellar electrokinetic chromatography (MEKC) was applied to the separation of a group of N-phenylpyrazole derivatives. Sodium dodecyl sulfate (SDS) as micellar system and 2-(N-cyclohexylamino)ethanesulfonic acid (CHES) as separation buffer (pH 10) were employed in the absence and presence of different percentages of medium chain alcohols (n-propanol or n-butanol). The separation of multicomponent mixtures of the solutes studied enabled the rapid determination of their retention factors which, in turn, allowed the study of the separation selectivity of compounds and the determination of their solute-micelle association constants (from the linear variation of the retention factors as a function of the total surfactant concentration in the separation buffer). Separation selectivity was studied according to the elution range and number of solutes separated in all the electrolyte solutions employed (45 micellar phases). The effect of the buffer concentration (0.05, 0.08 and 0.10 M), the alcohol nature (n-propanol or n-butanol) and the alcohol percentage (1, 3 or 5%) of the values obtained for the solute-micelle association constants was also studied. The best separation (12 solutes) was performed when a 0.08 M CHES buffer, pH 10, 0.02 M SDS modified by 5% n-butanol was used. The possibilities of using MEKC for evaluating the hydrophobicity of compounds was investigated through the study of the correlation between the logarithm of the retention factors of N-phenylpyrazole derivatives and their logarithm of the octanol-water distribution coefficients estimated by high performance liquid chromatography (HPLC).  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