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 791 毫秒
1.
Deoxyribonucleic acid (DNA) was electrochemically deposited on a carbon ionic liquid electrode to give a biosensor with excellent redox activity towards paraquat as shown by cyclic voltammetry and differential pulse voltammetry. Experimental conditions were optimized with respect to sensing paraquat by varying the electrochemical parameters, solution pH, and accumulation time of DNA. Under the optimized conditions, a linear relation exists between the reduction peak current and the concentration of paraquat in the range from 5?×?10?8 mol L?1 to 7?×?10?5 mol L?1, with a detection limit of 3.6?×?10?9 mol L?1. The utility of the method is illustrated by successful analysis of paraquat in spiked real water samples.
Figure
The DNA was electrodeposited onto the CILE under +1.5?V for 1200?s. The electrochemical behaviors of paraquat on the modified electrode had been studied by cyclic voltammetry and differential pulse voltammetry. Five ml phosphate buffer (pH 7.0) solution was added into an electrochemical cell (10?ml) and then paraquat was successfully added into the cell. The differential pulse voltammograms were recorded when swept from ?0.8?V to ?0.3?V. The peak currents at about ?0.63?V for paraquat were measured.  相似文献   

2.
The electrochemical impedance spectroscopy (EIS) has been used to study the interaction between paraquat and carbon modified by silver (Ag? CPE) and silver particles‐impregnated natural phosphate (Ag/NPh? CPE). This study was developed using spectrophotometry (UV? Vis) and infrared spectroscopy. The resulting interaction was controlled by adsorption at lower concentration (≤1.0×10?5 mol L?1) and by diffusion in the opposite case. Both electrodes are used to determining paraquat with a low detection limit (<1.0×10?12 mol L?1). The precision expressed as relative standard deviation RSD for the concentration level 1.0×10?5 mol L?1 of paraquat, (n=8) were 0.93 % and 1.1 % for Ag/NPh? CPE and Ag? CPE respectively.  相似文献   

3.
The voltammetric behavior of paraquat was investigated at hydroxyapatite‐modified carbon paste electrode HAP‐CPE in K2SO4. A method was developed for the detection of the trace of this herbicide, based on their redox reaction. The reduction peaks of paraquat were observed around ?0.70 V and ?1.00 V (vs. SCE) in square‐wave voltammetry. Experimental conditions were optimized by varying the accumulation time, apatite loading and measuring solution pH. Calibration plots were linear under the optimized parameters over the herbicide's concentration range 8–200×10?7 mol L?1, with a detection and quantification limits about 1.5×10?8 mol L?1 and 6.4 10?8 mol L?1, respectively.  相似文献   

4.
《Analytical letters》2012,45(4):694-705
A method for the determination of paraquat by cucurbit[7]uril (CB[7]) fluorescence quenching was developed. The assay was based on the reaction of the CB[7] with acridine orange. The fluorescence intensity of acridine orange regularly increased with the addition of CB[7]. However, while an appropriate amount of paraquat was added to the CB[7]- acridine orange system, the fluorescence intensity of the system was quenched which was employed to determine paraquat. Under the optimum conditions, a linear range of 3.0–800 nmol L?1 and a detection limit of 1.61 nmol L?1 for paraquat were obtained. The simple strategy reported here offers great practical potential for the determination of pesticide residues in agricultural products.  相似文献   

5.
The present work reports for the first time the determination of paraquat (PQ2+) by Differential Pulse Adsorptive Stripping Voltammetry (DPAdSV) using a carbon paste electrode modified (CPME) with biochar obtained from castor oil cake at different temperatures (200–600 °C). The best voltammetric response was verified using biochar yielded at 400 °C (CPME‐BC400). Linear dynamic range (LDR) for PQ2+ concentrations between 3.0×10?8 and 1.0×10?6 mol L?1 and a limit of detection (LOD) of 7.5×10?9 mol L?1 were verified. The method was successfully applied for PQ2+ quantification in spiked samples of natural water and coconut water.  相似文献   

6.
The anionic water-insoluble cyclodextrin polymer (polyCTR-β-CD) was crosslinked between β-cyclodextrin (β-CD) and citric acid (CTR) at 180?°C during 30?minutes to eliminate paraquat (PQ) from water. The reaction yield was equal to 70.2%, the ionic exchange capacity corresponded to 3.29?mmol·g?1 and the β-CD content was 0.29?mmol·g?1. Then, samples were characterized by SEM, ATR-FTIR, TGA, BET and stereoscopic microscope. Adsorption experiments were investigated with different factors such as pH of the solution, contact time, initial concentration of paraquat and adsorption temperature. The relevant pH was equal to 6.5 and the optimal contact time was 120?minutes to attain adsorption equilibrium. At 30?°C, the adsorption capacity was increased (9.4, 17.4 and 20.8?mg·g?1) when the initial concentration of paraquat was raised (25, 50 and 200?mg·L?1 respectively). Adsorption kinetics was appropriated to the pseudo-second-order model and adsorption isotherm was fitted to the Langmuir model. For thermochemistry parameters at different temperatures, the negative ΔG° showed a spontaneous adsorption process, the negative ΔH° indicated an exothermic process and the positive ΔS° exhibited an increase disorder. Finally, the reusability of the insoluble polymer was reached 78.3% after four regeneration cycles in methanol.  相似文献   

7.
《Electroanalysis》2006,18(23):2305-2313
The multiple square‐wave voltammetry (MSWV) allied to gold microelectrode (Au‐ME) was used to establish an electroanalytical procedure for the determination of the paraquat and diquat pesticides in river sediment samples. For both pesticides, two reduction peaks, at around ?0.70 V (peak 1) and around ?1.00 V vs. Ag/AgCl 3.00 mol L?1 (peak 2), with profile of the totally reversible redox process, were observed. The experimental and voltammetric conditions showed that the best conditions to reduce paraquat and diquat were a pH of 6.0, a frequency of 250 s?1, a scan increment 2 mV, a square‐wave amplitude of 50 mV and pulse number of 8 pulses of potential in each step of staircase of potential. Under such conditions, the detection limit of 0.044 μg L?1 (0.044 ppb) and 0.360 μg L?1 (0.360 ppb ) for peak 1 and peak 2 of paraquat and 0.159 μg L?1 (0.159 ppb) and 0.533 μg L?1 (0.533 ppb) for peak 1 and peak 2 of diquat, respectively, were obtained. These results are an order of magnitude of about two less than those obtained and published in the literature. Also, the electroanalytical procedure proposed was applied for the determination of adsorption isotherms of pesticides on river sediments samples collected from Mogi‐Guaçu River in Sao Paulo State, Brazil. The experimental data were fitted using the Langmuir and Freundlich isotherms models; and the results indicated low intensities of adsorption process of the pesticides in the samples employed with distribution coefficients (Kd) lower 5.0, and paraquat showed slightly higher affinity than diquat in the sediments. The increase in organic matter and organic carbon leads to an increase in the Kd values, and consequently an increase in the organic matter constant (KOM) organic carbon constant (KOC) values. All results demonstrated that isotherms “L” type in the Giles classification were obtained, indicating that sediments have a medium affinity for the pesticides, and no strong competition from the solvent used (in this case Na2SO4) for adsorption sites occurs.  相似文献   

8.
A bismuth‐film electrode (BiFE) ex situ electrochemically deposited onto a copper substrate has been presented for paraquat determination. The bismuth film was electrochemically deposited at an applied potential of ?0.18 V vs. Ag/AgCl (3.0 M KCl) for 200 s. The analytical curve was linear in the paraquat concentration range from 6.6×10?7 M to 4.8×10?5 M with a limit of detection of 9.3×10?8 M. The method presented satisfactory results at a confidence level of 95% and the performance was evaluated in water samples.  相似文献   

9.
The present work describes the first electrochemical method for quantifying paraquat herbicide poisoning in human saliva samples. Paraquat shows two couples of well‐defined peaks in aqueous solution using a boron doped diamond (BDD) electrode. By using square wave voltammetry (SWV) technique under optimum experimental conditions, a linear analytical curve was obtained for paraquat concentrations ranging from 0.800 to 167 µmol L?1, with a detection limit of 70 nmol L?1. This method was applied to quantify paraquat spikes in human saliva samples and in two different water samples (tap and river). The recovery values obtained ranged from 83.0 to 104 % and 99.1 to 105 %, respectively, which highlight the accuracy of the proposed method.  相似文献   

10.
《Analytical letters》2012,45(8):1491-1499
ABSTRACT

Glassy Carbon Electrodes coated with stearic acid provide an amperometric sensor for detection of paraquat, the active ingredient of the herbicide Gramoxone. The linear dynamic range of the sensor for Paraquat is 1.02 × 10?3 mol dm?3 to 1.02 × 10?2 mol dm?3 with the minimum detection limit 6.37 × 10?4 mol dm?3.  相似文献   

11.
The first three representatives of the new family of oxacyclophanes incorporating two 2,7‐dioxyfluorenone fragments, connected by [‐CH2CH2O‐]m spacers (m=2–4), have been synthesized. The yield of the smallest oxacyclophane (m=2) is considerably higher with respect to the larger ones (m=3 and m=4), which are formed in comparable yields. Molecular modeling and NMR spectra analysis of the model compounds suggest that an essential difference in oxacyclophanes yields is caused by formation of quasi‐cyclic intermediates, which are preorganized for macrocyclization owing to intramolecular π–π stacking interactions between the fluorenone units. The solid‐state structures of these oxacyclophanes exhibit intra‐ and intermolecular π–π stacking interactions that dictate their rectangular shape in the fluorenone backbone and crystal packing of the molecules with the parallel or T‐shape arrangement. The crystal packing in all cases is also sustained by weak C? H ??? O hydrogen bonds. FAB mass spectral analysis of mixtures of the larger oxacyclophanes (m=3 and m=4) and a paraquat moiety revealed peaks corresponding to the loss of one and two PF6? counterions from the 1:1 complexes formed. However, no signals were observed for complexes of the paraquat moiety with the smaller oxacyclophane (m=2). Computer molecular modeling of complexes revealed a pseudorotaxane‐like incorporation of the paraquat unit, sandwiched within a macrocyclic cavity between the almost parallel‐aligned fluorenone rings of the larger oxacyclophanes (m=3 and m=4). In contrast to this, only external complexes of the smallest oxacyclophane (m=2) with a paraquat unit have been found in the energy window of 10 kcal mol?1.  相似文献   

12.
A novel bis(m-phenylene)-26-crown-8-based cryptand has been synthesized. It has been used to prepare two 1:1 complexes with two paraquat derivatives with high association constants (6.5×105 and 4.0×105 M−1) in acetone. In the solid state the cryptand forms a 2:1 threaded structure with paraquat and an interesting supramolecular poly[2]pseudorotaxane threaded structure with a dihydroxyethyl-substituted paraquat derivative, respectively. It has been further used to prepare cryptand/paraquat derivative [2]rotaxanes efficiently by the immediate solvent evaporation method using easily available 3,5-dimethylphenyl groups as the stoppers.  相似文献   

13.
An electrochemical magneto immunosensor for the detection of low concentrations of paraquat (PQ) in food samples has been developed and its performance evaluated in a complex sample such as potato extracts. The immunosensor presented uses immunoreagents specifically developed for the recognition of paraquat, a magnetic graphite–epoxy composite (m-GEC) electrode and biofunctionalized magnetic micro-particles (PQ1-BSAMP) that allow reduction of the potential interferences caused by the matrix components. The amperometric signal is provided by an enzymatic probe prepared by covalently linking an enzyme to the specific antibodies (Ab198-cc-HRP). The use of hydroquinone, as mediator, allows recording of the signal at a low potential, which also contributes to reducing the background noise potentially caused by the sample matrix. The immunocomplexes formed on top of the modified MP are easily captured by the m-GEC, which acts simultaneously as transducer. PQ can be detected at concentrations as low as 0.18?±?0.09 μg L?1. Combined with an efficient extraction procedure, PQ residues can be directly detected and accurately quantified in potato extracts without additional clean-up or purification steps, with a limit of detection (90 % of the maximum signal) of 2.18?±?2.08 μg kg?1, far below the maximum residue level (20 μg kg?1) established by the EC. The immunosensor presented here is suitable for on-site analysis. Combined with the use of magnetic racks, multiple samples can be run simultaneously in a reasonable time.  相似文献   

14.
The photocatalytic degradation of paraquat (1,1-dimethyl-4,4′-bipyridylium dichloride) aqueous solutions in the presence of polycrystalline TiO2 Degussa P25 irradiated by near-UV light was investigated. The substrate and total organic carbon concentrations were monitored by UV spectroscopy and TOC measurements, respectively: the complete photocatalytic mineralization of paraquat (20 ppm) was achieved after ca. 3 h of irradiation by using 0.4 g l−1 of catalyst amount at natural pH (ca 5.8). On the contrary no significant photodegradation of paraquat was observed in the absence of TiO2 under similar experimental conditions. To evaluate the genotoxicity of paraquat and its intermediates produced during heterogeneous photocatalytic treatment, in vitro tests such as Ames test, with and without rat liver microsomal fractions (S9 mix), and micronucleus test, were used. Results obtained with Salmonella typhimurium (strain TA100) showed that paraquat and photocatalytic products were unable to induce gene mutations when photocatalysis was used in the presence of the optimum amount of TiO2, i.e. 0.4 g l−1, whereas an increase of revertants his+ per plate was observed after 300 min irradiation in the presence of very low amount of TiO2 (0.04 g l−1). The negative results from micronucleus test suggest that mutagenic, but non-clastogenic, late intermediates of paraquat photo-oxidation were formed when the photocatalytic runs of paraquat degradation were carried out by using 0.04 g l−1 of photocatalyst.  相似文献   

15.
The reactions of OH radical with Cl?, Br?, I?, and F? ions have been studied by entrapping the product radicals as polymer endgroup which have been detected and estimated by the sensitive dye partition technique. The rate constants of the reactions with Br?, Cl?, and F? ions have been determined to be 1.51 × 109, 1.32 × 109, and 0.92 × 109 L mol?1 s?1, respectively at 25°C and pH 1.00. Oxidation of I? ions liberates I, which inhibits the polymerization and the reaction could not be followed by polymer endgroup analysis. The observed order of reactivity Br? > Cl? > F? is in accordance with the electron affinities of the halide ions. The acidity of the reaction medium has a strong influence on the rate of reaction. With Br? ions, the rate constant of the reaction falls from 1.51 × 109 to 0.75 × 109 L mol?1 s?1 at 25°C as the pH is raised from 1.0 to 2.8. The method is simple and accurate and can be applied to study very reactive radicals.  相似文献   

16.

Paraquat dichloride commonly used as herbicide was determined by differential pulse cathodic stripping voltammetry technique. Experimental parameters, such as pH, accumulation time, accumulation potential and initial potential were optimized. In this analysis, paraquat dichloride exhibited a well-defined tworeduction peaks at ?0.35 and ?0.90 V in the pH range from 2.0 to 12.0. The 0.04 mol L–1 BR buffer at pH 2.0 was found a suitable medium for electroanalytical determination of the paraquat dichloride. Interfering ions effect was not significant. Linear calibration plots for standard solutions of paraquat dichloride were obtained in the range of 0.25 to 1.75 × 10–6 mol L–1. Detection limit was 3.66 × 10–8 mol L–1. The optimized parameters were effectively applied for the determination of commercial paraquat dichloride and in artificial samples. Artificial samples were prepared by spiking paraquat dichloride into tap water and drinking water dispenser samples. The recovery value was 90.5% in drinking water dispenser samples and 91.7% in tap water samples at the concentration range of 1.00 × 10–6 to 1.75 × 10–6 mol L–1.

  相似文献   

17.
A method for the determination of the herbicides diquat and paraquat in water was developed using liquid chromatography-(electrospray ionization) mass spectrometry [LC-(ESI)MS]. The analytes were isolated on an ENVI-8 DSK solid phase extraction (SPE) disk and eluted with 5-M trifluoroacetic acid (TFA). The eluate was evaporated to dryness and the analytes were redissolved in the mobile phase (7% methanol/93% water/25-mM TFA). The extract was analyzed by liquid chromatography (C1 column) with postcolumn addition of propionic acid/methanol followed by (ESI)MS. Diquat was detected using the [M2+ ? H+] ion (M2+ = dication) at m/z 183, whereas paraquat was detected using the mono-trifluoroacetate ion pair [M2+/?OOCCF3] at m/z 299. Quantitation was done by isotope dilution mass spectrometry using d 4-diquat and d 8-paraquat and the corresponding ions [M2+ ? D+] and [M2+/?OOCCF3] at m/z 186 and m/z 307, respectively. Detection limits of 0. 1 and 0. 2 µg/L, respectively (based on the dications), were adequate to meet the Ontario Drinking Water Objectives of 70 and 10 µg/L, respectively, and the Ontario Provincial Water Quality Objective for diquat of 0. 5 µg/L. Precision and accuracy were 14% and 6% for diquat and 12% and 3% for paraquat.  相似文献   

18.
Two molecular shuttles/switches—a slow one and a fast one—in the shape of amphiphilic, bistable [2]rotaxanes have been synthesized and characterized. Both [2]rotaxanes contain a hydrophobic, tetraarylmethane and a hydrophilic, dendritic stopper. They are comprised of two π‐electron‐rich stations—a monopyrrolotetrathiafulvalene unit and a 1,5‐dioxynaphthalene moiety—which can act as recognition sites for the tetracationic cyclophane, cyclobis(paraquat‐p‐phenylene), to reside around. In addition, a model [2]rotaxane, incorporating only a monopyrrolotetrathiafulvalene unit in the rod section of the amphiphilic dumbbell component and cyclobis(paraquat‐p‐phenylene) as the ring component, has been investigated. The dumbbell‐shaped components were constructed using conventional synthetic methodologies to assemble 1) the hydrophobic, tetraarylmethane stopper and 2) the hydrophilic, dendritic stopper. Next, 3) the hydrophobic stopper was fused to the 1,5‐dioxynaphthalene moiety and/or the monopyrrolotetrathiafulvalene unit by appropriate alkylations, followed by 4) attachment of the hydrophilic stopper, once again by alkylation to give the dumbbell‐shaped compounds. Finally, 5) the [2]rotaxanes were self‐assembled by using the dumbbells as templates for the formation of the encircling cyclobis(paraquat‐p‐phenylene) tetracations. The two [2]rotaxanes differ in their arrangement of the π‐electron‐rich units, one in which the SMe group of the monopyrrolotetrathiafulvalene unit points toward the 1,5‐dioxynaphthalene moiety ( 2 ?4 PF6) and another in which it points away from the 1,5‐dioxynaphthalene moiety ( 3 ?4 PF6). This seemingly small difference in the orientation of the monopyrrolotetrathiafulvalene unit leads to profound changes in the physical properties of these rotaxanes. The bistable [2]rotaxanes were both isolated as brown solids. 1H NMR and UV‐visible spectroscopy, and electrochemical investigations, reveal the presence of both possible translational isomers at ambient temperature. As a consequence of the existence of both possible translational isomers in these bistable [2]rotaxanes, they exhibit a complex electrochemical behavior, which is further complicated by the presence of folded conformations wherein the monopyrrolotetrathiafulvalene unit is involved in an “alongside” interaction with the tetracationic cyclophane. In the molecular shuttle/switch 2 ?4 PF6 a “knob”, in the shape of the SMe group, is situated between the monopyrrolotetrathiafulvalene and the 1,5‐dioxynaphthalene recognition sites, making it possible to isolate both translational isomers ( 2 ?4 PF6?GREEN and 2 ?4 PF6?RED) and to investigate the kinetics of the shuttling of the cyclobis(paraquat‐p‐phenylene) tetracation between the two recognition sites. The shuttling processes, which are accompanied by clearly detectable color changes, can be followed by 1H NMR and UV‐visible spectroscopy, allowing the rate constants and energies of activation for the translation of the cyclobis(paraquat‐p‐phenylene) tetracations between the two recognition sites to be determined. In the molecular shuttle/switch 3 ?4 PF6, there is no “knob” situated between the 1,5‐dioxynaphthalene and the monopyrrolotetrathiafulvalene recognition sites, resulting in a considerably faster shuttling of the cyclobis(paraquat‐p‐phenylene) tetracation between these two sites, making the separation of the two possible translational isomers of 3 ?4 PF6 impractical. However, the shuttling of the cyclobis(paraquat‐p‐phenylene) tetracation can be followed by dynamic 1H NMR spectroscopy. At low temperatures, the major translational isomer is 3 ?4 PF6?RED, while 3 ?4 PF6?GREEN is the major isomer at higher temperature. In the bistable [2]rotaxanes shuttling of the cyclobis(paraquat‐p‐phenylene) tetracations can be driven by electrochemical oxidation of the monopyrrolotetrathiafulvalene unit. In complexes in which one of the two dumbbell stoppers is missing, electrochemical oxidation causes dethreading.  相似文献   

19.
A fast and sensitive photokinetic method for the determination of paraquat (MV2+) (1?27×10?5M) is described, based on the rate of photoreduction of MV2+ by EDTA, sensitized by acridine yellow in the absence of oxygen. The rate of photoreduction, which is a linear function of the concentration of MV2+ is monitored polarographically by recording the limiting current of p- benzoquinone, which is reduced by the radical monocation MV generated in the photochemical reaction. The results obtained by the application of the fixed-time, fixed-concentration change and initial-rate kinetic methods are evaluated. An alternative method for monitoring the rate of the process is by measuring the time necessary for the total reduction of p-benzoquinone. The end-point is detected with two platinum electrodes at an applied voltage of 100 mV. The procedure has been successfully applied to the determination of paraquat in commercial herbicides, waters, and flowers and in spiked soils and blood sera.  相似文献   

20.
《Analytical letters》2012,45(11):1917-1928
Abstract

A fast and simple electrogenerated unstable reagent chemiluminescence system for flow injection analysis of etamsylate is described, based on a decrease in the chemiluminescence intensity from the luminol-hypochlorite system. The response is linear to the concentration of etamsylate in the range from 1x109 to 8x 10?9. The detection limit is 6 x 10?10 g ml?1. The relative standard deviation is 3.1% at 2 x 10?9 g ml?1 (n=7). The proposed method is suitable for automatic and continuous analysis and has been successfully tested for determination of etamsylate in pharmaceutical formulations.

  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