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 234 毫秒
1.
An environment‐sensitive fluorophore can change its maximum emission wavelength (λem), fluorescence quantum yield (Φf), and fluorescence lifetime in response to the surrounding environment. We have developed two new intramolecular charge‐transfer‐type environment‐sensitive fluorophores, DBThD‐IA and DBSeD‐IA, in which the oxygen atom of a well‐established 2,1,3‐benzoxadiazole environment‐sensitive fluorophore, DBD‐IA, has been replaced by a sulfur and selenium atom, respectively. DBThD‐IA is highly fluorescent in n‐hexane (Φf=0.81, λem=537 nm) with excitation at 449 nm, but is almost nonfluorescent in water (Φf=0.037, λem=616 nm), similarly to DBD‐IA (Φf=0.91, λem=520 nm in n‐hexane; Φf=0.027, λem=616 nm in water). A similar variation in fluorescence properties was also observed for DBSeD‐IA (Φf=0.24, λem=591 nm in n‐hexane; Φf=0.0046, λem=672 nm in water). An intensive study of the solvent effects on the fluorescence properties of these fluorophores revealed that both the polarity of the environment and hydrogen bonding with solvent molecules accelerate the nonradiative relaxation of the excited fluorophores. Time‐resolved optoacoustic and phosphorescence measurements clarified that both intersystem crossing and internal conversion are involved in the nonradiative relaxation processes of DBThD‐IA and DBSeD‐IA. In addition, DBThD‐IA exhibits a 10‐fold higher photostability in aqueous solution than the original fluorophore DBD‐IA, which allowed us to create a new robust molecular nanogel thermometer for intracellular thermometry.  相似文献   

2.
A new nickel(II) complex, [Ni(o-van-L-met)(phen)(CH3OH)] (o-van-L-met = Schiff base derived from o-vanillin and l-methionine, phen = 1,10–phenanthroline), has been synthesized and characterized by elemental analyses, IR spectra, and single-crystal X-ray diffraction. The crystal structure shows nickel is six-coordinate in a distorted octahedral geometry. In this crystal, molecules form a 2-D plane structure via hydrogen bonds and π–π interactions. The interaction of the complex with calf thymus DNA (CT-DNA) was investigated by absorption, fluorescence, circular dichroism (CD), spectroscopies, and viscosity measurement. The complex binds to CT-DNA in an intercalative mode with a binding constant of (4.7 ± 0.5) × 104 M?1. The interaction of the complex with bovine serum albumin (BSA) was also studied by the multispectroscopic methods. Results illustrated that the nickel(II) complex can effectively quench the intrinsic fluorescence of BSA via a static quenching mechanism and cause conformational changes. The binding constant Kb was (6.3 ± 1.6) × 104 M?1 and the binding site number n was 0.96 ± 0.04; its bind site was located within subunit IIA of BSA.  相似文献   

3.
The fluorescence detection of di-phosphonic acid and mono-phosphonic acid derivatives using the anthracene-based diamidine 1 has been investigated. The diamidine 1 forms 1:1 and 1:2 complexes with the di-phosphonic acid and mono-phosphonic acid derivatives, respectively, and showed a blue fluorescence (λem = 432–442 nm) in a DMSO solution. The formation of amidinium-phosphonate (complex formation) and dissociated amidinum (λem = 468 nm as a broad band) were distinguished by the difference in the fluorescence wavelength, and confirmed by DOSY NMR spectroscopy and TD-DFT calculations. The formation of a 1:2 complex with diamidine 1 and methylphosphonic acid having additional intermolecular hydrogen-bonding between the methylphosphonic acids is proposed.  相似文献   

4.
Chrysin, 7-hydroxyflavone, and quercetin were studied for their affinities with human serum albumin (HSA) in the presence and absence of Fe2+ and Co2+. The fluorescence intensities of HSA decrease remarkably with increasing concentration of the tested flavonoids. Chrysin resulted in a blue-shift of the emission line λ em of HSA from 336 to 330 nm whereas quercetin showed an obvious red-shift of λ em from 336 to 347 nm. However, the extents of the λ em shifts induced by flavonoids in the presence of mental ions are much bigger than those of the corresponding systems in the absence of mental ions. Fe2+ and Co2+ increased the quenching constants of the tested flavonoids for HSA by 12.4–48.1 and 15.0–66.7 %, respectively. The affinities of 7-hydroxyflavone, chrysin and quercetin for HSA increased by about 6.42, 7.38 and 0.62 %, respectively, in the presence of Fe2+. Co2+ increased the affinities of 7-hydroxyflavone, chrysin, and quercetin for HSA about 8.43, 7.86 and 11.73 %, respectively.  相似文献   

5.
The fluorescent peptidocalixarenes, 5,11,17,23-tetra-tert-butyl-25,26,27,28-tetrakis(O-methyl-l-tryptophanylcarbonylmethoxy)calix[4]arene (1) and 5,11,17,23-tetra-tert-butyl-25,27-di(O-methyl)-26,28-bis(O-methyl-l-tryptophanylcarbonylmethoxy)calix[4]arene (2), were prepared by introducing tryptophan subunits at a lower calixarene rim. Coordination abilities of 1 and 2 towards Eu(III) and alkali metal cations were studied by spectrophotometric, spectrofluorimetric, conductometric and potentiometric titrations in acetonitrile at 25°C. Rather strong complexation was observed for smaller alkali metal cations Li+ and Na+ (log K Li1 >6, log K Li2 >6, log K Na1  = 8.25, log K Na2  = 6.94), and moderate for K+ (log K K1  = 5.09, log K K2  = 4.09). Larger Rb+ and Cs+ cations did not fit in the ion binding site of 1 so no complexation was detected, whereas the more flexible ligand 2 accommodated Rb+ cation (log K Rb2  = 3.44). The fluorescence of 1ex = 280 nm, λem = 340 nm) was remarkably quenched by Eu(III). Stability constant of 1:1 (Eu3+:1) complex determined spectrofluorimetrically amounted to log K Eu1  = 6.16.  相似文献   

6.

A validated, selective and sensitive pre-column derivatization chromatographic method for determination of some aminoglycoside antibiotics, namely, amikacin (AMK), gentamicin and neomycin sulphates, has been developed. Determination was obtained by pre-column reaction of the samples with 7-fluoro-4-nitrobenz-2-oxa-1,3-diazole and separating the corresponding derivatives on a reversed phase LC column (ultrasphere C18, 250 × 4.6 mm) and fluorescence detection (λ ex 465 nm, λ em 531 nm). Linear quantitative response curve was generated over a concentration range of 0.05–10 μg mL−1 with a correlation coefficient of more than 0.999 with all studied compounds. The accuracy and precisions were satisfactory for determination of all drugs in dosage forms within- and between-run assay. Furthermore, the method was extended to the determination of AMK in spiked human plasma. The method is specific for the intact drugs, and can be adopted in the presence of co-formulated drugs.

  相似文献   

7.
It is very important to clarify the mechanisms of the interaction between components of organisms. In this report, the interaction between bovine serum albumin (BSA) and ionic polysaccharides were discussed. The fluorescence spectrum of tryptophan (Trp) in BSA changed as its conformation changed. On adding polysaccharide containing sulfonic acid residues (sulfonic‐type) at pH 7.4, a remarkable blue shift of the emission maximum (λem) of Trp was observed. This blue shift was decreased by adding NaCl. In contrast, polysaccharide containing carboxylic acid residues (carboxy‐type) scarcely changed Trp fluorescence at the same pH. At a pH lower than the isoelectric point (PI = 4.7–4.9) of BSA, a remarkable blue shift was observed not only by adding sulfonic‐type polysaccharide but also by adding carboxy‐type polysaccharide. Moreover, using the energy transfer method, in the pH region higher than the PI of BSA, carboxylic‐type polysaccharides may interact relatively weakly with the binding site II of BSA, but sulfonic‐type ones can selectively interact with binding site I weakly and with binding site II strongly. And in the pH region lower than the PI of BSA, carboxylic‐type polysaccharides interact with binding site II strongly. On the other hand, sulfonic‐type polysaccharides are bound to both binding site I and binding site II very strongly. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

8.
The nitridophosphates AEP8N14 (AE=Ca, Sr, Ba) were synthesized at 4–5 GPa and 1050–1150 °C applying a 1000 t press with multianvil apparatus, following the azide route. The crystal structures of CaP8N14 and SrP8N14 are isotypic. The space group Cmcm was confirmed by powder X-ray diffraction. The structure of BaP8N14 (space group Amm2) was elucidated by a combination of transmission electron microscopy and diffraction of microfocused synchrotron radiation. Phase purity was confirmed by Rietveld refinement. IR spectra are consistent with the structure models and the chemical compositions were confirmed by X-ray spectroscopy. Luminescence properties of Eu2+-doped samples were investigated upon excitation with UV to blue light. CaP8N14 (λem=470 nm; fwhm=1380 cm−1) and SrP8N14 (λem=440 nm; fwhm=1350 cm−1) can be classified as the first ultra-narrow-band blue-emitting Eu2+-doped nitridophosphates. BaP8N14 shows a notably broader blue emission (λem=417/457 nm; fwhm=2075/3550 cm−1).  相似文献   

9.
We demonstrate a reversible shape‐morphing with concurrent fluorescence switching in the nanomaterials which are complexed with cucurbit[7]uril (CB[7]) in water. The cyanostilbene derivative alone forms ribbon‐like two‐dimensional (2D) nanocrystals with bright yellow excimeric emission in water (λem=540 nm, ΦF=42 %). Upon CB[7] addition, however, the ribbon‐like 2D nanocrystals immediately transform to spherical nanoparticles with significant fluorescence quenching and blue‐shifting (λem=490 nm, ΦF=1 %) through the supramolecular complexation of the cyanostilbene and CB[7]. Based on this reversible fluorescence switching and shape morphing, we could demonstrate a novel strategy of turn‐on fluorescence sensing of spermine and also monitoring of lysine decarboxylase activity.  相似文献   

10.
Four d10 transition-metal coordination complexes 1–4 (1: [Ag2(TPT)2(TPTH)2]; 2: [Cu6(TPT)6]·2DMF; 3: [Cd(TPT)2(TPTH)]·CH3CH2OH, 4: [Zn(TPT)2]n) have been constructed from a newly designed heterocyclic thioamide ligand, TPTH (TPTH = 4-(thiophen-2-yl)-pyrimidine-2-thiol). All complexes have been structurally elucidated by single crystal X-ray diffraction analyses. Except for 4, which shows a one-dimensional (1-D) chain structure, 1–3 are all discrete coordination complexes featuring dinuclear, hexanuclear and mononuclear entities, respectively. Their photophysical properties have been evaluated in the solid state at room temperature by UV–vis diffuse reflectance and luminescence spectra. Among them, 2 exhibits a strong red luminescence (λem = 699 nm) with a remarkable red-shift of the maximum emission compared to that of the TPTH ligand (λem = 536 nm). The red emission observed with 2 is ascribed to a LMCT (ligand-to-metal charge transfer) transition which agrees with the DFT calculations.  相似文献   

11.
A new chromium(III) complex, [CrCl(naph-gly)phen]?H2O (naph-gly = Schiff base derived from 2-hydroxy-1-naphthaldehyde and glycine, phen = 1,10-phenanthroline), has been synthesized and characterized by elemental analysis, electrospray ionization mass spectroscopy, FT-IR, and X-ray single-crystal diffraction. The chromium(III) complex belongs to the trigonal crystal system, P3(1) space group with crystallographic data: a = b = 1.97017(16) nm, c = 1.02991(7) nm, α?=?β?=?90°, γ =120°, V = 3.4621(5) nm3, Dc = 1.476 g?cm?3, Z = 6, F(0 0 0)?=?1578, R1 = 0.0508, wR2 = 0.0907. There are two independent molecules in the crystallographic asymmetric unit of the chromium(III) complex. Each CrIII is six-coordinate to form an octahedral geometry. In the crystal, a 3-D structure is formed through intermolecular hydrogen bonds. The calf thymus DNA (CT-DNA)- and bovine serum albumin (BSA)-binding properties of the complex have been studied by UV absorption, fluorescence, and circular dichroism (CD) spectroscopy. Results indicate that the chromium(III) complex binds to CT-DNA in an intercalative mode, and it can bind to BSA and cause conformational changes of BSA.  相似文献   

12.
The interaction of Schiff base ligand A and its three metal complexes [A‐Fe(II), A‐Cu(II), and A‐Zn(II)] with bovine serum albumin (BSA) was investigated using a tryptophan fluorescence quenching method. The Schiff base ligand A and its three metal complexes all showed quenching of BSA fluorescence in a Tris‐HCl buffer. Quenching constants were determined for quenching BSA by the Schiff base ligand A and its metal complexes in a Tris‐HCl buffer (pH=7.4) at different temperatures. The experimental results show that the dynamic quenching constant (KSV) was increased with increasing temperature, whereas the association constant (K) was decreased with the increase of temperature. The thermodynamic parameters ΔH, ΔG and ΔS at different temperatures were calculated. The ionic strength of the Tris‐HCl buffer had a great influence on the wavelength of maximum emission of BSA. Under low ionic strength, the emission spectra of BSA influenced by A‐Zn(II) had a small blue shift. Compared to A‐Zn(II), the emission spectra of BSA in the presence of the Schiff base ligand A and A‐Cu(II) had no significant λem shift. At high ionic strength, the emission spectra of BSA upon addition of the Schiff base A, A‐Fe(II), and A‐Zn(II) all had a red shift, but the emission spectra of BSA had λem shift neither at low ionic strength, nor at high ionic strength in the presence of A‐Cu(II). Furthermore, the temperature did not affect the λem shift of BSA emission spectra.  相似文献   

13.
A validated, selective and sensitive pre-column derivatization chromatographic method for determination of some aminoglycoside antibiotics, namely, amikacin (AMK), gentamicin and neomycin sulphates, has been developed. Determination was obtained by pre-column reaction of the samples with 7-fluoro-4-nitrobenz-2-oxa-1,3-diazole and separating the corresponding derivatives on a reversed phase LC column (ultrasphere C18, 250 × 4.6 mm) and fluorescence detection (λ ex 465 nm, λ em 531 nm). Linear quantitative response curve was generated over a concentration range of 0.05–10 μg mL?1 with a correlation coefficient of more than 0.999 with all studied compounds. The accuracy and precisions were satisfactory for determination of all drugs in dosage forms within- and between-run assay. Furthermore, the method was extended to the determination of AMK in spiked human plasma. The method is specific for the intact drugs, and can be adopted in the presence of co-formulated drugs.  相似文献   

14.
Two non-natural amino acids with fluorescent urea side-chains were prepared from Fmoc-protected aspartic and glutamic acids. In acetonitrile solution, the emission of the Asp derivative is strongly quenched by HCO3 or H2PO4 (K ≥ 104 M? 1) but not by less-basic Cl or NO3. Solutions containing excess bicarbonate ion appear peach-colored, with λabs at 394 and 495 nm ascribed to the anion complex and urea-deprotonated sensor, respectively. Corresponding fluorescence bands are observed at 475 and 579 nm. Dihydrogenphosphate is not sufficiently basic to remove H+ from the ground state of the fluorophore. However, deprotonation of the excited state occurs in the presence of>1 equiv of H2PO4em = 578 nm). According to 1H NMR in DMSO-d6, recognition of H2PO4 occurs at the urea N–H groups and the amino acid backbone N–H. DFT techniques further predict that the backbone C = O group accepts an H-bond from the anion. The Glu derivative has lower affinity for anions; the additional CH2 group in its side-chain apparently sets the backbone N–H and C = O too far from the urea to contribute significantly to binding. To demonstrate suitability for standard Fmoc-based solid-phase peptide synthesis, the Asp derivative was incorporated into a 12-residue peptide.  相似文献   

15.
Two polymorphs emitting near‐infrared ( 1 R form: α phase, λem=702 nm, Φf=0.41) and orange‐red fluorescence ( 1 O form: β phase, λem=618 nm, Φf=0.05) were synthesized by finely controlling the crystallization conditions of compound 1 , a structurally simple excited‐state intramolecular proton transfer (ESIPT)‐active molecule. Multicolor amplified spontaneous emissions (ASEs) were realized, for the first time, based on these polymorphs. Notably, the 1 O crystal underwent heating‐induced phase transformation from the β phase to the α form in a single‐crystal to single‐crystal (SCSC) manner accompanied with an unprecedented ASE changing. The ASE behavior of polymorphs 1 R , 1 O as well as the ASE changing during SCSC was investigated. The feasibility of multicolor lasing based on the present organic polymorphs was confirmed, which may provide a new development strategy for organic laser science and technology.  相似文献   

16.
DNA and BSA binding properties of mononuclear Co (II) and Cu (II) complexes containing letrozole [M(Le)4Cl2]·(H2O)](Le=[4,4-(1H-1,2,4-triazol-1-ylmethylene)bisbenzonitrile] have been investigated under physiological conditions. The interaction ability of the two complexes with native calf thymus DNA(CT-DNA) has been monitored as a function of the metal complex-DNA molar ratio by UV–Vis absorption spectrophotometry, fluorescence spectroscopy, circular dichroism(CD) and thermal denaturation studies. The intrinsic binding constants, Kb, of complexes 1 and 2 with CT-DNA, obtained from UV–Vis absorption studies, were 3.15 ± 0.02 × 104 and 4.37 ± 0.02 × 104 M?1, respectively. The addition of the complexes to CT-DNA (1:2) leads to an increase in the melting temperature of DNA up to around 4 °C, which has revealed that complexes could interact with DNA through intercalation mode. Fluorimetric studies have been performed using methylene blue (MB) as a fluorescence probe and competitive studies have shown the ability of the complexes to displace the DNA-bound MB, suggesting competition with MB. To explore the potential biological value of the complexes, the binding interaction between Co (II) and Cu (II) complexes and bovine serum albumin (BSA) has also been studied by fluorescence spectroscopy. The results indicate that the reaction between the complexes and BSA is a static quenching procedure. The site marker displacement experiment has suggested the location of the complexes binding to BSA at Sudlow’s site I in subdomain IIA. Finally, MTT assay studies have shown that the bioactive complexes exert significantly high selective dose-dependent cytotoxicity against a panel of cancer cell lines including MCF-7, JURKAT, SKOV3 and U87.  相似文献   

17.
A new copper(II) complex, [Cu(Bphen)(Leu)(NO3)(H2O)] (Bphen = 4,7-diphenyl-1,10-phenanthroline, leu = L-leucine), has been synthesized and characterized by IR spectroscopy, CHN analysis, and single-crystal X-ray diffraction techniques. The CT-DNA binding properties of the complex have been investigated by both absorption and emission spectroscopy. The binding parameters for the fluorescence Scatchard plot were also determined. Further, the interaction of the complex with bovine serum albumin (BSA) has been investigated using absorption and emission spectroscopy. The thermodynamic parameters, free energy change (ΔG), enthalpy change (ΔH), and entropy change (ΔS), were calculated by the van’t Hoff equation and discussed. The distance between BSA and the complex has been obtained according to fluorescence resonance energy transfer. Conformational changes of BSA have been observed from synchronous fluorescence. Antioxidant and radical scavenging activities of the complex were determined by various in vitro assays such as 1,1-diphenyl-2-picryl-hydrazyl free radicals (DPPH˙), 2,2′-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) radicals (ABTS˙+), and reducing ability determination by H2O2 scavenging methods.  相似文献   

18.
《Analytical letters》2012,45(1):104-118
Abstract

In this study a spectrofluorimetric method was developed to determine drug active compound in the tablets for albendazole (ABZ). For this aim, fluorescence spectra of albendazolee drug active compound in various solvents were taken, it was determined that the most suitable solvent was chloroform and excitation (λex) and emission (λem) wavelength in a row was 360 nm and 440 nm in this solvent. Calibration graphics were drawn and shown linear at 0–2.5 mg/l concentration range (R2=0.9993). Albendazole quantity in the andazol tablets was determined from directly calibration graphs and with standard addition method. Results obtained with developed spectrofluorimetric method were compared with standard USP method and it was found no difference between two methods within 95% confidence limits. It was determined that proposed method was easy and highly accuracy method to be able to use in routine albendazole analyze for the quality control.  相似文献   

19.
Three cadmium(II) coordination polymers [Cd(NA)2(H2O)2]n ( 1 ), {[Cd(NA)(phen)(NO3)]·(H2O)1/2}n ( 2 ), {[Cd(NA)(CH3C6H4COO)(H2O)2]·(CH3C6H4COOH)}n ( 3 ) (HNA = nicotinic acid, phen = 1, 10‐phenanthroline) have been synthesized by hydrothermal method. Their single‐crystal structures were determined by X‐ray diffractometry. The absorption, excitation and emission spectra were investigated and all the complexes emit strong fluorescence: λemmax = 544 nm (λex = 492 nm), 1 ; λemmax = 466 nm (λex = 393 nm), 2 ; λemmax = 430 nm (λex = 313 nm), 3 . At room temperature in the solid state the fluorescence lifetimes of the complexes were investigated and the relationships between the spectra were discussed as well as the connections of luminescence and crystal structures.  相似文献   

20.
The binding interaction of a Schiff base compound containing a 1,2,4‐triazole ring [4‐(4‐chlorobenzyl‐ideneamino)‐5‐methyl‐1,2,4‐triazole‐3‐thiol, CTT] with bovine serum albumin (BSA) was studied by spectroscopy methods including fluorescence and circular dichroism spectrum under simulative physiological conditions. Fluorescence investigation revealed that the fluorescence quenching of BSA was induced by the formation of a relative stable CTT‐BSA complex. The corresponding binding constants (Ka) between CTT and BSA at three different temperatures were calculated according to the modified Stern‐Volmer equation. The enthalpy change (ΔH) and entropy change (ΔS) were calculated to be −15.78 kJ·mol−1 and 49.23 J·mol−1·K−1, respectively, which suggested that hydrophobic forces and hydrogen bond played major roles in stabilizing the CTT‐BSA complex. Site marker competitive experiments indicated that the binding of CTT to BSA primarily took place in sub‐domain IIIA (site II) of BSA. The binding distance (r) between CTT and the tryptophan residue of BSA was obtained to be 4.3 nm based on F?rster theory of non‐radioactive energy transfer. The conformational investigation revealed that the presence of CTT decreased the α‐helix content of BSA (from 58.62% to 54.66%) and induced the slight unfolding of the polypeptides of protein, which confirmed some micro‐environmental and conformational changes of BSA molecules.  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