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.
The effect of titanium dioxide (TiO2) nanoparticles (NPs) on photophysical characteristics of 2,5‐di[(E)‐1‐(4‐dimethylaminophenyl) methylidine]‐1‐cyclopentanone (2,5‐DMAPMC) and 2,5‐di[(E)‐1‐(4‐diethylaminophenyl)methylidine]‐1‐cyclopentanone (2,5‐DEAPMC) ketocyanine dyes has been studied using absorption, steady‐state and time‐resolved fluorescence spectroscopy. The magnitudes of association constants determined based on modified absorption spectrum of dyes due to the presence of TiO2 NPs indicate the interaction of TiO2 NPs with dye molecules. The quenching of fluorescence intensity of dyes by TiO2 NPs is observed and it follows linear Stern‐Volmer (S‐V) equation. The magnitude of quenching rate parameter suggests the involvement of static quenching mechanism. The involvement of electron transfer process in reducing fluorescence intensity of dyes has been discussed. Also, varying influence of TiO2 NPs on two dyes is explained based on the presence of different alkyl substituent in two dyes.  相似文献   

2.
The mechanism of charge generation in solid‐state dye‐sensitized solar cells using triarylamine‐substituted perylene monoimide dyes is studied by vis‐NIR broadband pump‐probe transient absorption spectroscopy. The experiments demonstrate that photoinduced electron injection into the TiO2 can only occur in regions where Li+, from the commonly used Li‐TFSI additive salt, is present on the TiO2 surface. Incomplete surface coverage by Li+ means that some dye excitons cannot inject their electron into the TiO2. However it is observed in the solar cell structure that some of the dye excitons that cannot directly inject an electron still contribute to free charge generation by the previously hypothesized reductive quenching mechanism (hole transfer to the solid‐state hole transporter followed by electron injection from the dye anion into the TiO2). The contribution of reductive quenching to the quantum efficiency of charge generation is significant, raising it from 68% to over 80%. Optimization of this reductive quenching pathway could be exploited to maintain high quantum efficiency in dyes with greater NIR absorption to achieve overall enhancements in device performance. It is demonstrated that broadband NIR transient spectroscopy is necessary to obtain population kinetics in these systems, as strong Stark effects distort the population kinetics in the visible region.  相似文献   

3.
The present work investigated the inorganic carbon (Ci) uptake, fluorescence quenching and photo‐inhibition of the edible cyanobacterium Ge‐Xian‐Mi (Nostoc) to obtain an insight into the role of CO2 concentrating mechanism (CCM) operation in alleviating photo‐inhibition. Ge‐Xian‐Mi used HCO3 in addition to CO2 for its photosynthesis and oxygen evolution was greater than the theoretical rates of CO2 production derived from uncatalysed dehydration of HCO3. Multiple transporters for CO2 and HCO3 operated in air‐grown Ge‐Xian‐Mi. Na+‐dependent HCO3 transport was the primary mode of active Ci uptake and contributed 53–62% of net photosynthetic activity at 250 µmol L?1 KHCO3 and pH 8.0. However, the CO2‐uptake systems and Na+‐independent HCO3 transport played minor roles in Ge‐Xian‐Mi and supported, respectively, 39 and 8% of net photosynthetic activity. The steady‐state fluorescence decreased and the photochemical quenching increased in response to the transport‐mediated accumulation of intracellular Ci. Inorganic carbon transport was a major factor in facilitating quenching during the initial stage and the initial rate of fluorescence quenching in the presence of iodoacetamide, an inhibitor of CO2 fixation, was 88% of control. Both the initial rate and extent of fluorescence quenching increased with increasing external dissolved inorganic carbon (DIC) and saturated at higher than 200 µmol L?1 HCO3. The operation of the CCM in Ge‐Xian‐Mi served as a means of diminishing photodynamic damage by dissipating excess light energy and higher external DIC in the range of 100–10000 µmol L?1 KHCO3 was associated with more severe photo‐inhibition under strong irradiance.  相似文献   

4.
1,3‐Diphenyl‐5‐(9‐anthryl)‐2‐pyrazoline and 1,3‐diphenyl‐5‐(9‐anthryl)‐1H‐pyrazole with an anthryl chromophore were synthesized and characterized using 1H NMR, 13C NMR, FT‐IR, mass spectrometry and elemental analysis. Their optical properties were characterized by UV–vis absorption and fluorescence spectroscopy. It was observed that the absorption and fluorescence spectra of the two compounds showed a red shift with respect to that of anthracene. Pyrazole exhibited high fluorescent quantum yields (Φf = 0.90 in toluene) while pyrazoline showed nearly no fluorescence in solution. The significant fluorescence divergence of the two similar compounds was investigated theoretically through density functional theory (DFT) calculations. The energetically lowest‐lying state S1 in the pyrazoline exhibited both characteristics of locally excited and electron‐transfer states that resulted in the fluorescence quenching of anthryl chromophore whereas the S1 state in the pyrazole corresponded to an optically allowed state that led to high fluorescence quantum yields in solutions. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

5.
Light-induced changes in the fluorescence of the pH-indicating dyes pyranine or 5-(and 6-)carboxy-2, 7-dichlorofluorescein (CDCF) which had been fed to leaves were examined to monitor cellular pH changes. After short-term feeding of pyranine (pK 7.3) to leaves of Amaranthus caudatus L., a NAD-malic-enzyme-type C4 plant, vascular bundles and surrounding cells became fluorescent. Fluorescence emission from mesophyll cells required longer feeding times. In CO2-free air, pyranine fluorescence increased much more on illumination after mesophyll cells had become fluorescent than when only the vascular bundles and the bundle sheath of Amaranthus leaves had been stained. After short feeding times and in the absence of actinic illumination, CO2 decreased pyranine fluorescence very slowly in Amaranthus and rapidly in C3 leaves. After prolonged feeding times, the extent of the light-dependent increase in pyranine fluorescence was several times greater in different C4 plants than in C3 species. The kinetics of the fluorescence changes were also remarkably different in C3 and C4 plants. Carbon dioxide (500 l · l–1) suppressed the light-induced increase in pyranine fluorescence more in C4 than in C3 leaves. Light-dependent changes in light scattering, which are indicative of chloroplast energization, and in 410-nm transmission, which indicate chloroplast movement, differed kinetically from those of the changes in pyranine fluorescence. Available evidence indicated that light-dependent changes in pyranine fluorescence did not originate from the apoplast of leaf cells. Microscopic observation led to the conclusion that, after prolonged feeding times or prolonged incubation, changes in pyranine fluorescence emitted from C4 leaves reflect pH changes mainly in the cytosol of mesophyll cells. A transient acidification reaction indicated by quenching of pyranine fluorescence in the dark-light transient and not observed in C3 species is attributed to the carboxylation of phosphoenolpyruvate. After short feeding times and in the absence of actinic illumination, CO2 (250 l l–1) decreased pyranine fluorescence very slowly in Amaranthus and more rapidly in C3 leaves. After prolonged feeding times, both the rate and the extent of CO2-dependent quenching of pyranine fluorescence increased, but the increase was insufficient to indicate the presence of highly active carbonic anhydrase in the compartment from which pyranine fluorescence was emitted. In contrast to pyranine, CDCF (pK 4.8) did not increase but rather decreased its fluorescence on illumination of an Amaranthus leaf, indicating acidification of an acidic compartment, most probably the vacuole of green leaf cells. The pattern of the acidification reaction was similar in C4 and C3 leaves. The remarkably large extent of the light-dependent increase in pyranine fluorescence from leaves of C4 species and its slow kinetics are proposed to be caused by an alkalization of the cytosol which in the absence of CO2 is larger in the mesophyll than in the bundle sheath. It gives rise to deprotonation of dye originally located in the mesophyll and, in addition, of dye which diffuses from the bundle sheath into the mesophyll following a pH gradient. Implications of slow diffusional transport of pyranine and CO2 between mesophyll and bundle-sheath cells and the fast metabolite transport required in C4 photosynthesis are discussed.Abbreviations CDCF 5-(and 6-)carboxy-2,7-dichlorofluorescein - DHAP dihydroxyacetone phosphate - PGA 3-phosphoglycerate This work was supported by the Sonderforschungsbereiche 176 and 251 of the University of Würzburg and by the Gottfried-Wilhelm-Leibniz Program of the Deutsche Forschungsgemeinschaft. A.S.R. was the recipient of a fellowship of the Alexander-von-Humboldt Foundation. We are grateful to Mrs. S. Neimanis for cooperation.  相似文献   

6.
A novel ligand‐assisted assembly approach is demonstrated for the synthesis of thermally stable and large‐pore ordered mesoporous titanium dioxide with a highly crystalline framework by using diblock copolymer poly(ethylene oxide)‐b‐polystyrene (PEO‐b‐PS) as a template and titanium isopropoxide (TIPO) as a precursor. Small‐angle X‐ray scattering, X‐ray diffraction (XRD), transmission electron microscopy (TEM), high‐resolution scanning electron microscopy, and N2‐sorption measurements indicate that the obtained TiO2 materials possess an ordered primary cubic mesostructure with large, uniform pore diameters of about 16.0 nm, and high Brunauer–Emmett–Teller surface areas of ~112 m2 g?1, as well as high thermal stability (~700 °C). High resolution TEM and wide‐angle XRD measurements clearly illustrate the high crystallinity of the mesoporous titania with an anatase structure in the pore walls. It is worth mentioning that, in this process, in addition to tetrahydrofuran as a solvent, acetylacetone was employed as a coordination agent to avoid rapid hydrolysis of the titanium precursor. Additionally, stepped evaporation and heating processes were adopted to control the condensation rate and facilitate the assembly of the ordered mesostructure, and ensure the formation of fully polycrystalline anatase titania frameworks without collapse of the mesostructure. By employing the obtained mesoporous and crystallized TiO2 as the photoanode in a dye‐sensitized solar cell, a high power‐conversion efficiency (5.45%) can be achieved in combination with the N719 dye, which shows that this mesoprous titania is a great potential candidate as a catalyst support for photonic‐conversion applications.  相似文献   

7.
The interaction of glucose‐derived carbon quantum dots (CQDs) with silver (Ag) and gold (Au) nanoparticles (NPs) was explored by fluorescence spectroscopy. Both metal NPs cause an efficient quenching of CQD fluorescence, which is likely due to the energy transfer process between CQDs as donors and metal NPs as acceptors. The Stern–Volmer plots were evaluated and corresponding quenching constants were found to be 1.9 × 1010 and 2.2 × 108 M?1 for AgNPs and AuNPs, respectively. The analytical applicability of these systems was demonstrated for turn‐on fluorescence detection of the anti‐cancer drug, 6‐thioguanine. Because the CQD–AgNP system had much higher sensitivity than the CQD–AuNP system, we used it as a selective fluorescence probe in a turn‐on assay of 6‐thioguanine. Under optimum conditions, the calibration graph was linear from 0.03 to 1.0 μM with a detection limit of 0.01 μM. The developed method was applied to the analysis of human plasma samples with satisfactory results.  相似文献   

8.
To develop conducting organic polymers (COPs) as luminescent sensors for determination of toxic heavy metals, a new benzene sulfonic acid‐doped polypyrrole (PPy‐BSA) thin film was electrochemically prepared by cyclic voltammetry (CV) on flexible indium tin oxide (ITO) electrode in aqueous solution. PPy‐BSA film was characterized by FTIR spectrometry, X‐ray photoelectron spectroscopy (XPS) and scanning electron microscopy (SEM). The optical properties of PPy‐BSA were investigated by ultraviolet (UV)‐visible absorption and fluorescence spectrometry in dimethylsulfoxide (DMSO) diluted solutions. PPy‐BSA fluorescence spectra were strongly quenched upon increasing copper(II) ion (Cu2+) and lead(II) ion (Pb2+) concentrations in aqueous medium, and linear Stern–Volmer relationships were obtained, which indicated the existence of a main dynamic fluorescence quenching mechanism. BSA‐PPy sensor showed a high sensitivity for detection of both metallic ions, Cu2+ and Pb2+, with very low limit of detection values of 3.1 and 18.0 nM, respectively. The proposed quenching‐fluorimetric sensor might be applied to the determination of traces of toxic heavy metallic ions in water samples.  相似文献   

9.
The molecular interactions between salicylic acid (SA) and proflavin hemisulfate (PF) were investigated using fluorescence and UV–VIS absorption spectroscopy in an aqueous micellar environment. Changes in the absorption spectra of SA in the presence of PF indicate a ground state interaction between salicylate and proflavine hemisulfate ions to form a complex. The excitation bands of SA monitored at its emission wavelength reveal a red spectral shift of 8390.54 and 2037.75 cm‐1 when compared with absorption bands. The intensity of both excitation bands decreased in the presence of increasing amounts of PF. The absence of excitation bands of PF rules out the possibility of its direct excitation and suggests energy transfer from excited SA to PF, resulting in quenching of the SA fluorescence. The fluorescence quenching results were found to fit the well‐known Stern–Volmer (S–V) relation. S–V plots at different temperatures were used to further evaluate thermodynamic parameters such as ?G, ?H and ΔS. The thermodynamic and kinetic data obtained from the quenching results were used to investigate the possible mechanism of binding, the nature of the binding force and the distance between SA and PF molecules. The linear relation between SA fluorescence quenching and PF concentration used to develop an analytical method for the determination of PF from Lorexane (a veterinary cream) using a fluorescence quenching method. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

10.
A responsive spectrofluorometric method was developed for the determination of sitagliptin phosphate using l -tyrosine as a fluorescence probe. The fluorescence intensity of l -tyrosine was quenched with sitagliptin phosphate. The fluorescence intensity was recorded at 307 nm using a 272 nm excitation wavelength. The calibration plot between fluorescence intensity and the concentration of drug was linear in the range of 0.1 to 2.0 mM with a good correlation value of 0.997. The limit of detection and quantification were established to be 3.7 × 10−4 and 1.23 × 10−3 mM, respectively. Commonly used excipients did not interfere with sitagliptin phosphate measurement. The proposed method was used to measure the sitagliptin phosphate in its standard type, dosage form, and biological samples. The percent recovery ranged from 97.41–103.36%. The static quenching was shown to be responsible for quenching as indicated by the Stern–Volmer plot. The method was validated using ICH guidelines and profitably applied for the content uniformity test, resulting in a high percent recovery and small relative standard deviation. The proposed approach is effortless, susceptible, selective, economic, and provides a high precision and accuracy, and can be used to determine sitagliptin phosphate in the pharmaceutical industry.  相似文献   

11.
Quenching effects of bergenin, based on the electrochemiluminescence (ECL) of the tris(2,2′‐bipyridyl)‐ruthenium(II) (Ru(bpy)32+)/tri‐n‐propylamine (TPrA) system in aqueous solution, is been described. The quenching behavior can be observed with a 100‐fold excess of bergenin over Ru(bpy)32+. In the presence of 0.1 m TPrA, the Stern–Volmer constant (KSV) of the ECL quenching is as high as 1.16 × 104 M?1 for bergenin. The logarithmic plot of the inhibited ECL versus logarithmic plot of the concentration of bergenin was linear over the range 3.0 × 10?6–1.0 × 10?4 mol/L. The corresponding limit of detection was 6.0 × 10?7 mol/L for bergenin (S/N = 3). In the mechanism of quenching it is believed that the competition of the active free radicals between Ru(bpy)32+/TPrA and bergenin was the key factor for the ECL inhibition of the system. Photoluminescence, cyclic voltammetry, coupled with bulk electrolysis, supports the supposition mechanism of the Ru(bpy)32+/TPrA–bergenin system. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

12.
Jing Li  Shuhua Han 《Luminescence》2017,32(8):1448-1455
A novel chemosensor TrisRh based on tris(2‐aminoethyl)amine and rhodamine 6G is designed and synthesized as a fluorescence turn‐on probe for Co2+ ions that is paramagnetic with a property of quenching fluorescence. Rhodamine spirolactam forms are nonfluorescent, whereas, ring‐opening of corresponding spirocyclic induced by Co2+ results in strong fluorescence emission. Upon the addition of Co2+ ions, TrisRh can display significant enhancements in absorbance and fluorescence intensity as well as evident colorific transformation, which can be perceived by the naked eye. The association stoichiometry of TrisRh to Co2+ ions was inferred to be 1:1 through Job's plot and electrospray ionization mass spectrometry analysis. The binding model was speculated from Fourier transform infrared spectra and 1H–nuclear magnetic resonance technologies. Significantly, the limit of detection was determined to be as low as 1.22 nmol. Furthermore, TrisRh can exhibit robust anti‐jamming ability against other interference metal ions.  相似文献   

13.
In this study, a high fluorescence sensitivity and selectivity, molecularly imprinted nanofluorescent polymer sensor (MIP@SiO2@QDs) was prepared using a reverse microemulsion method. 2,4,6‐Trichlorophenol (2,4,6‐TCP) was detected using fluorescence quenching. Tetraethyl orthosilicate (TEOS), quantum dots (QDs) and 3‐aminopropyltriethoxysilane (APTS) were used as cross‐linker, signal sources and functional monomer respectively. The sensor (MIP@SiO2@QDs) and the non‐imprinted polymer sensor (NIP@SiO2@QDs) were characterized using infra‐red (IR) analysis, X‐ray diffraction (XRD), transmission electron microscopy (TEM) and scanning electron microscopy (SEM). The selectivity of MIP@SiO2@QDs was examined by comparing 2,4,6‐TCP with other similar functional substances including 2,4‐dichlorophenol (2,4‐DCP), 2,6‐dichlorophenol (2,6‐DCP) and 4‐chlorophenol (4‐CP). Results showed that MIP@SiO2@QDs had better selectivity for 2,4,6‐TCP than the other compounds. Fluorescence quenching efficiency displayed a good linear response at the 2,4,6‐TCP concentration range 5–1000 μmol/L. The limit of detection (LOD) was 0.9 μmol/L (3σ, n = 9). This method was equally applicable for testing actual samples with a recovery rate of 98.0–105.8%. The sensor had advantages of simple pretreatment, good sensitivity and selectivity, and wide linear range and could be applied for the rapid detection of 2,4,6‐TCP in actual samples.  相似文献   

14.
We estimated the relative florescence quantum yield (Φ) of 8‐methoxy‐3‐[1‐(4,5‐dicarbomethoxy‐1,2,3‐triazoloacetyl)]coumarin [8MDTC] using a single‐point method with quinine sulfate in 0.1 M of sulfuric acid used as a standard reference. The fluorescence lifetimes, radiative and non‐radiative decay rate constants are calculated. Relative quantum yields were found to be less in the non‐polar solvents, indicating that the solute exhibits less fluorescence in a non‐polar environment. The fluorescence quenching of [8MDTC] by aniline was studied at room temperature by examining the steady state in five different solvents in order to explore various possible quenching mechanisms. The experimental results show a positive deviation in Stern–Volmer plots in all solvents. Ground state complex and sphere of action static quenching models were used to interpret the results. Many quenching rate parameters were calculated using these models. The values of these parameters suggest that the sphere of action static quenching model agrees well with the experimental results. Further, a finite sink approximation model was used to check whether these bimolecular reactions were diffusion limited or not. The values of the distance parameter R′ and the diffusion coefficient D were determined and are compared with the values of the encounter distance R and diffusion coefficient D calculated using the Stokes–Einstein equation. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

15.
A sipholane triterpenoid, named sipholenone A, with anti‐cancer properties was isolated from the Red Sea sponge Siphonochalina siphonella and characterized by proton and carbon‐13 nuclear magnetic resonance (1H NMR and 13C NMR) spectroscopies. The goal of this study was to visualize the binding of this triterpenoid with human serum albumin (HSA) and to determine its binding site on the biomacromolecule. The interaction was visualized using fluorescence quenching, synchronous fluorescence, far‐ and near‐UV circular dichroism (CD), UV–visible and Fourier transform‐infrared (FT‐IR) spectroscopies. UV–visible spectroscopy indicated the formation of a ground‐state complex as a result of the interaction. Sipholenone A quenches the fluorescence of HSA via a static quenching mechanism. A small blue shift in the fluorescence quenching profiles suggested the involvement of hydrophobic forces in the interaction. Sipholenone A binding takes place at site I of subdomain II A with a 1:1 binding ratio, as revealed by displacement binding studies using warfarin, ibuprofen and digitoxin. Far‐UV CD and FT‐IR studies showed that the binding of sipholenone A to HSA also had a small effect on the protein's secondary structure with a slight decrease in the α‐helical content. Several thermodynamic parameters were calculated, along with Forster's radiative energy transfer analysis.  相似文献   

16.
The light-dependent quenching of 9-aminoacridine fluorescence was used to monitor the state of the transthylakoid proton gradient in illuminated intact chloroplasts in the presence or absence of external electron acceptors. The absence of appreciable light-dependent fluorescence quenching under anaerobic conditions indicated inhibition of coupled electron transport in the absence of external electron acceptors. Oxygen relieved this inhibition. However, when DCMU inhibited excessive reduction of the plastoquinone pool in the absence of oxygen, coupled cyclic electron transport supported the formation of a transthylakoid proton gradient even under anaerobiosis. This proton gradient collapsed in the presence of oxygen. Under aerobic conditions, and when KCN inhibited ribulose bisphosphate carboxylase and ascorbate peroxidase, fluorescence quenching indicated the formation of a transthylakoid proton gradient which was larger with oxygen in the Mehler reaction as electron acceptor than with methylviologen at similar rates of linear electron transport. Apparently, cyclic electron transport occured simultaneously with linear electron transport, when oxygen was available as electron acceptor, but not when methylviologen accepted electrons from Photosystem I. The ratio of cyclic to linear electron transport could be increased by low concentrations of DCMU. This shows that even under aerobic conditions cyclic electron transport is limited in isolated intact chloroplasts by excessive reduction of electron carriers. In fact, P700 in the reaction center of Photosystem I remained reduced in illuminated isolated chloroplasts under conditions which resulted in extensive oxidation of P700 in leaves. This shows that regulation of Photosystem II activity is less effective in isolated chloroplasts than in leaves. Assuming that a Q-cycle supports a H+/e ratio of 3 during slow linear electron transport, vectorial proton transport coupled to Photosystem I-dependent cyclic electron flow could be calculated. The highest calculated rate of Photosystem I-dependent proton transport, which was not yet light-saturated, was 330 mol protons (mg chlorophyll h)–1 in intact chloroplasts. If H+/e is not three but two proton transfer is not 330 but 220 mol (mg Chl H)–1. Differences in the regulation of cyclic electron transport in isolated chloroplasts and in leaves are discussed.  相似文献   

17.
Zinc toxicity on photosynthetic activity in cells of Synechocystis aquatilis f. aquatilis Sauvageau was investigated by monitoring Hill activity and fluorescence. The oxygen‐evolving activity decreased to about 80% of the initial value after exposure to 0.1 mM ZnSO4 for 1 h. The PSII activity was inhibited by 40% in the presence of zinc concentrations ranging from 0.5 to 5.0 mM, suggesting that the metal effect is limited by zinc uptake. The fluorescence capacity (Fmax–F/Fmax) decreased from 0.57 to 0.35 and 0.20 in Zn‐treated cells for 15 and 60 min, respectively, thus providing evidence for rapid inactivation of electron transport at PSII. Zinc treatment promoted a rapid increase in PSII fluorescence that was counteracted by addition of 1,4‐benzoquinone, indicating that electron transfer at the reducing side of the PSII reaction center is arrested by zinc. Furthermore, a decline in the fluorescence yield could be observed after 1 h of zinc treatment as well as when Zn‐treated cells were excited in presence of 3‐(3′,4′‐dichlorophenyl)‐1,1‐dimethylurea. Under these conditions, zinc did not affect energy transfer from phycobilisomes to PSII, and the gradual quenching of PSII fluorescence may be due to a decrease in electron flow on the donor side of PSII. However, the 20% increase in the minimal fluorescence intensity (Fo) in parallel to the absence of changes in the maximal fluorescence intensity (Fmax), observed in the first hour of zinc treatment, could also suggest a metal‐induced decline in the energy transfer from PSII‐chl a antenna to the PSII reaction center.  相似文献   

18.
The ineraction between riboflavin (RBF) and tryptophan (Trp) was investigated using fluorescence spectroscopy and UV–vis absorption spectroscopy under physiological conditions. The fluorescence of Trp was quenched by RBF via dynamic quenching, which was analyzed using the Stern–Volmer relation. The value of the Forster distance R0 (2.31 nm) was obtained according to the Forster's theory of nonradiative energy transfer. Under physiological conditions, a linear relationship could be established between the quenched fluorescence intensity of Trp and the concentration of RBF in the range of 5.8 × 10‐7–2.0 × 10‐5 mol/L. The detection limit was 1.8 × 10‐7 mol/L. The method was successfully applied to determine riboflavin concentrations in pharmaceutical samples. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

19.
A sensitive fluorogenic probe 1 for biothiols was developed based on the Michael addition reaction. The probe 1 was readily synthesized via the reaction of 2‐(4′‐hydroxyphenyl) benzimidazole (HPBI) with acryloyl chloride and shows weak fluorescence emission. Upon mixing with biothiols, the fluorescence of 1 is significantly enhanced due to the conjugate addition of thiols to the α,β‐unsaturated carbonyl moiety, thus eliminating the photoinduced electron transfer (PET) quenching of the fluorophore by the intramolecular carbon–carbon double bond. Cysteine (Cys) was selected as the representative thiol in the spectral experiment. A good linear relationship was obtained from 1.0 to 30.0 µmol L?1 for Cys and the detection limit was 0.17 µmol L?1. Furthermore, probe 1 was highly selective for biothiols without the interference of some biologically relevant analytes and has been applied to detecting biothiols in human urine samples. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

20.
We report on a simple and sensitive sulfur and nitrogen co‐doped carbon quantum dot (S,N‐CQD)‐based chemiluminescence (CL) sensor for the determination of indomethacin. S,N‐CQDs were prepared by a hydrothermal method and characterized by fluorescence spectra, Fourier transform infrared spectroscopy and transmission electron microscopy. To obtain the best CL system for determination of indomethacin, the reaction of S,N‐CQDs with some common oxidants was studied. Among the tested systems, the S,N‐CQD–KMnO4 reaction showed the highest sensitivity for the detection of indomethacin. Under optimum conditions, the calibration plot was linear over a concentration range of 0.1–1.5 mg L?1, with a limit of detection (3σ) of 65 μg L?1. The method was applied to the determination of indomethacin in environmental and biological samples with satisfactory results.  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