首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The interaction of the coronary vasodilator dipyridamole with biological systems, protein and membranes has been studied through optical absorption and fluorescence spectroscopies. Using the analysis of the spectra and fluorescence intensity of dipyridamole (DIP) in solution, the interaction of this compound with the transport protein albumin (BSA) and with a model of cell membranes, namely micelles of lysophosphatidylcholine (L-PC), was investigated. Measurements were performed at pH 5.0 and pH 7.0 where the molecule of DIP is fully protonated and partially protonated, respectively. The quenching of fluorescence with nitroxide-stable radicals 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) and 4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPOL) as well as with acrylamide and iodide allowed the localization of the drug in the polar interface of micelles. Quenching by acrylamide and iodide in L-PC micelles demonstrated the effect of micelle protonation which increased the accessibility of iodide to the chromophore. An effective association constant was obtained both at pH 7.0 (7.5 x 10(3) M-1) and pH 5.0 (2.5 x 10(3) M-1) and a very good agreement with the proposed binding model was observed. The quantum yields of fluorescence data agree very well with the fluorescence lifetimes. The measurement of lifetimes was important to understand the kinetic data obtained from Stern-Volmer plots both of radical, acrylamide and iodide quenching of fluorescence. It was observed that, in the presence of micelles, the kq value increased for TEMPO while decreased for TEMPOL. This result, together with the vanishing solubility of DIP in saturated hydrocarbons and the preferential partition of TEMPO in micelles, suggested the localization of DIP in the polar micellar interface. This is also supported by the enhanced iodide quenching at pH 5.0, constancy of acrylamide quenching in the range of pH 7.0-5.0 and the partition of TEMPO and TEMPOL in SDS micelles. The association constant of DIP to BSA was also estimated both at pH 7.0 (2 x 10(4) M-1) and pH 5.0 (4 x 10(3) M-1). Quenching studies with nitroxide radicals, acrylamide and iodide also suggested the binding of the drug to a hydrophobic region of the protein. At pH 5.0, the protein undergo a conformational change which leads to a loosening of the overall structure so that the accessibility of the nitroxide radicals for DIP is increased at this pH. The differences in kq values at pH 7.0 and pH 5.0 suggested that at pH 7.0 the chromophore is protected in the protein site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Quenching of anthracycline fluorescence by a series of spin-labeled fatty acids was used to probe the transverse location of the drug in phosphatidylcholine bilayers in the form of small unilamellar vesicles. Stern-Volmer plots of the quenching data indicate that the fluorophore moiety of the anthracycline is intercalated into the hydrocarbon region of the bilayer, with deeper penetration observed in fluid-phase than in solid-phase vesicles. 31P-NMR parameters (T1 and nuclear Overhauser enhancement (NOE] are unaffected by the presence of drug, consistent with a binding site removed from the interfacial region. Comparison of intensity (F0/F) plots with lifetime (tau 0/tau) data shows that the predominant mechanism of anthracycline quenching by membrane-bound nitroxides is static. Since the membrane-bound drug is also accessible to quenching by I-, the binding site in the membrane must create a channel which is accessible to solvent. Two other fluorescent probes, 12-(9-anthroyloxy)stearate (12-AS) and diphenylhexatriene (DPH), were employed to confirm the results obtained with the anthracyclines, giving quenching data representative of their location in the bilayer.  相似文献   

3.
The goal of the present study is to elucidate the mechanism of quercetin on modulating Naja naja atra phospholipase A2 (PLA2) activities. Sphingomyelin inhibited PLA2 enzymatic activity and membrane-damaging activity against egg yolk phosphatidylcholine (EYPC), while cholesterol and quercetin abrogated the sphingomeyelin inhibitory effect. Quercetin incorporation led to a reduction in PLA2 enzymatic activity and membrane-damaging activity toward EYPC/sphingomyelin/cholesterol vesicles. Both cholesterol and quercetin increased detergent resistance and reduced membrane fluidity of EYPC/sphingomyelin vesicles. Quercetin reduced detergent insolubility but increased ordered lipid packing of EYPC/sphingomyelin/cholesterol vesicles. Acrylamide quenching studies and trinitrophenylation of Lys residues revealed that quercetin altered the membrane-bound mode of PLA2 differently upon absorption onto the membrane bilayers of different lipid compositions. However, 8-anilinonaphthalene sulphonate-binding assay revealed that quercetin marginally affected the interaction between active site of PLA2 with phospholipid vesicles. Collectively, our data indicate that membrane-inserted quercetin modulates PLA2 interfacial activity and membrane-damaging activity via its effects on membrane structure and membrane-bound mode of PLA2.  相似文献   

4.
Interactions between the fluorophors diphenylhexatriene or gramicidin A′ and lipids are examined using a spin-labeled phosphatidylcholine as a fluorescence quenching probe. It is found that in phospholipid vesicles of mixed lipid composition at temperatures where phospholipids are completely in the liquid crystal phase, several different species of phosphatidylcholines are randomly distributed around the fluorophors. In vesicles of mixed lipid composition which can undergo thermally induced phase separations, the fluorescence quenching observed at lower temperatures reflects a non-random distribution of lipids around each fluorophor. This observation is explained in terms of the partition of fluorophor between a spin-labeled lipid-rich liquid crystal phase, and a spin-labeled lipiddepleted gel phase. Gramicidin A′ strongly favors partition into the liquid crystal phase, whereas diphenylhexatriene partitions about equally between the two lipid phases. A method is described utilizing fluorescence quenching for the calculation of the partition coefficient for a fluorophor. The partition coefficients so calculated are shown to be in good agreement with previously reported values derived from other methods. It is also shown that Ca2+-induced lipid phase separations can be monitored by fluorescence quenching.  相似文献   

5.
The distribution of free fatty acids at equilibrium after incubation of small sonicated unilamellar vesicles (SUV) with large unilamellar vesicles (LUV) of different lipid composition has been determined. Stearic acid (SA) and oleic acid (OA) showed similar preferences for SUV and LUV of egg yolk phosphatidylcholine (EYPC). Both ionized and protonated forms of the free fatty acids (FFAs) behaved similarly with respect to the equilibrium distribution between EYPC of different size. The charge of the vesicles was found, however, to be important, since both FFAs in their ionized form preferentially associated to vesicles of phosphatidylcholine (PC) as compared with vesicles of phosphatidylglycerol (PC). While SA preferred membranes in the gel state, OA showed preference for the membrane in fluid state. The insertion of both OA and SA in phosphatidylethanolamine (PE)/phosphatidylcholine vesicles is less favourable than in vesicles of pure PC. All these data suggest that membrane lipid content may play a role in determining the distribution of free fatty acids among the membranes of a cell.  相似文献   

6.
The interaction of the tetramisole derivative (+-)-5,6-dihydro-6-phenyl-imidazo[2,1-b]thiazole and a number of its 2-n-alkyl homologues (-ethyl through -n-pentyl and -n-heptyl) with large unilamellar phosphatidylcholine/phosphatidylethanolamine/dipalmitoylphosphatidic acid (2:1:0.06, w/w) vesicles was studied by means of steady-state fluorescence quenching using 8-(2-anthryl)octanoic acid as membrane probe. Linear Stern-Volmer plots were obtained for each derivative, indicating dynamic quenching. The slopes of the plots decreased with increasing liposomal concentration. For four short-chain homologues (-H, -ethyl, -n-propyl and -n-butyl), the respective membrane partition coefficients Kp and bimolecular quenching rate constants kq were determined from the plots of the reciprocal of the apparent quenching rate constant (kappq)-1 against the lipid volume fraction alpha L of the liposomes. The partition coefficients increased with increasing chain-length of the tetramisoles. A linear relationship was found between the free energy of partitioning and the number of methylene units of the homologues (-delta G degrees per methylene group = 1.6 +/- 0.1 kJ mol-1). For the n-pentyl and n-heptyl derivatives, the fluorescence quenching technique did not allow one to determine their membrane partition coefficients. Analysis of the fluorescence intensity measurements with Scatchard plots gave further evidence for the partitioning nature of the tetramisole derivatives' association with the liposomal membranes.  相似文献   

7.
To elucidate the contribution of phospholipase A2 (PLA2) activity of notexin to its ability to perturb membranes, comparative studies on the interaction of notexin and guanidinated notexin (Gu-notexin) with egg yolk phosphatidylcholine (EYPC), EYPC/egg yolk sphingomyelin (EYSM) and EYPC/EYSM/cholesterol vesicles were conducted. EYSM notably reduced the membrane-damaging activity of notexin against EYPC vesicles, but had an insignificant influence on that of Gu-notexin. Unlike the effects noted with notexin, inactivation of PLA2 activity by EDTA led to a reduction in the ability of Gu-notexin to induce EYPC/EYSM vesicle leakage and to increase Gu-notexin-induced membrane permeability of EYPC/EYSM/cholesterol vesicles. The geometrical arrangement of notexin and Gu-notexin in contact with either EYPC/EYSM vesicles or EYPC/EYSM/cholesterol vesicles differed. Moreover, global conformation of notexin and Gu-notexin differed in either Ca2+-bound or metal-free states. These results indicate that notexin and Gu-notexin could induce membrane permeability without the involvement of PLA2 activity, and suggest that guanidination alters the membrane-bound mode of notexin on damaging phospholipid vesicles containing sphingomyelin and cholesterol.  相似文献   

8.
To examine physical-chemical aspects of bile salt-phospholipid interactions that could contribute to preferential phosphatidylcholine (PC) secretion into bile, we have compared transitions between vesicles and micelles in model systems containing taurocholate (TC) and either egg-yolk PC (EYPC), egg-yolk sphingomyelin (EYSM), buttermilk SM (BMSM) or dipalmitoyl PC (DPPC). Phase transitions from micelles to vesicles were observed at 4-fold dilution of serially diluted EYPC/TC systems, but not earlier than at 16-fold dilution of SM/TC or DPPC/TC systems, indicating lower concentrations of the detergent required for micellization in the case of SM or DPPC. Cryo-transmission electron microscopy of phase transitions initiated by addition of TC to phospholipid vesicles revealed extremely long SM-containing intermediate structures, but shorter EYPC-containing intermediate structures. Again, larger amounts of bile salt were required to induce phase transitions in the case of EYPC compared to SM. Sizes of TC-phospholipid micelles increased progressively upon increasing phospholipid contents in the rank order: DPPC-TC相似文献   

9.
The tryptophan intrinsic fluorescence of mitochondrial complex III reconstituted in phosphatidylcholine bilayers was examined at different temperatures. Absorption and emission maxima occur at 277 and 332 nm, irrespective of temperature or lipid:protein ratio even if there are indications (from fluorescence quenching) of protein conformational changes as a function of lipid:protein ratio. Low values of Trp fluorescence quantum yield in complex III (0.008-0.010) are probably due to the neighborhood of the heme groups. The temperature-dependent decrease of fluorescence intensity is nonlinear; the corresponding Arrhenius plots show "breaks" or discontinuities that could be interpreted as thermally dependent changes in protein conformation. However, no temperature-dependent changes in fluorescence quenching have been observed that may be related to protein conformational changes. In addition, Arrhenius plots of the fluorescence intensity of simple molecules, such as Trp or 1-anilino-8-naphthalene sulfonate in the presence of aqueous phospholipid dispersions, also show breaks in the same temperature range. Stern-Volmer plots of acrylamide and iodide quenching were also nonlinear, indicating large differences in quenching constants for the various tryptophanyl residues. The quenching results also suggest that, at high lipid:protein ratios, the microviscosity of the protein matrix is higher than that in lipid-poor systems. Comparison of quenching efficiencies of iodide and acrylamide suggest that no significant fraction of the fluorophores occurs in the neighborhood of charged residues.  相似文献   

10.
Melittin free of phospholipase A2 was prepared. In the absence of salt this highly pure protein starts to aggregate in solution at a protein concentration of Cp greater than 10(-3) M. In high salt solution (2 M) aggregation starts at Cp greater than 10(-6) M. This was determined from the blue shift of the intrinsic fluorescence of the protein. Reinvestigation of the quenching behaviour clearly shows that self-aggregation cannot be deduced from quenching experiments using nitrate or 2,2,6,6-tetramethylpiperidine-1-oxyl as quencher. The incorporation of melittin into phosphatidylcholine bilayer vesicles was studied by fluorescence quenching and by energy-transfer experiments using 2- and 6-anthroyloxypalmitic acid as acceptor and peptide tryptophan as donor. Incorporation of melittin into small unilamellar vesicles was found to be reduced below the lipid phase transition temperature, Tt, whereas it incorporates and distributes more randomly above Tt. Cooling the temperature below Tt after incubation at T greater than Tt leads to a deeper incorporation of the peptide into the lipid bilayer due to electrostatic interaction between the lipid phosphate groups and the positively charged amino acids. This stabilizing effect is lost above Tt and melittin is extruded to the polar phase. Quenching experiments support this finding. EPR measurements clearly demonstrate that even in the presence of high amounts of melittin up to 10 mol% with respect to the lipid broadening of the phase transition curves was only observed with fatty acid spin labels, where the doxyl group is localized near the bilayer surface. The order degree of the inner part of the bilayer remains almost unchanged even in the presence of high melittin content.  相似文献   

11.
T G Burke  T R Tritton 《Biochemistry》1985,24(21):5972-5980
We have exploited the intrinsic fluorescence properties of the anthracycline antitumor antibiotics to study the dependence on drug structure of relative drug location and dynamics when the anthracyclines were bound to sonicated dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylcholine (DPPC) vesicles at 27.5 degrees C. Iodide quenching experiments at constant ionic strength were used to evaluate the relative accessibilities of the bound fluorophores to membrane-impermeable iodide. Iodide was found to quench the fluorescence of anthracyclines in free solution by both static and dynamic mechanisms, whereas quenching of membrane-bound fluorophores was predominantly due to the dynamic mechanism. Modified Stern-Volmer plots of anthracyclines bound to fluid-phase DMPC bilayers were linear, and the biomolecular rate constant (kq) values ranged from 0.6 X 10(9) to 1.3 X 10(9) M-1 s-1. Modified Stern-Volmer plots of anthracyclines bound to solid-phase DPPC bilayers were curved, indicative of a heterogeneous-bound drug population. A strong correlation between drug hydrophobicity and penetration of the fluorophore into the bilayer was observed for the daunosamine-containing anthracyclines. Steady-state fluorescence anisotropy measurements under iodide quenching conditions were used to investigate the diffusive motions of anthracyclines in isotropic solvent and in fluid-phase DMPC bilayers. Anthracycline derivatives free in solution exhibited limiting anisotropy (alpha infinity) values which decayed to zero at times long compared to the excited-state lifetime, in contrast to anthracyclines bound to fluid-phase DMPC bilayers, which showed nonzero alpha infinity values. Steady-state anisotropies of membrane-bound anthracyclines were found to be governed principally by alpha infinity and not by the mean rotational rate (R).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The conformational states of Ca2+-ATPase in sarcoplasmic reticulum (SR) vesicles with or without a thousand-fold transmembrane Ca2+ gradient have been studied by fluorescence spectroscopy and fluorescence quenching. In consequence of the establishment of the transmembrane Ca2+ gradient, the steady-state fluorescence results revealed a reproducible 8% decrease in the intrinsic fluorescence while time-resolved fluorescence measurements showed that 13 tryptophan residues in SR · Ca2+-ATPase could be divided into three groups. The fluorescence lifetime of one of these groups increased from 5.5 ns to 5.95 ns in the presence of a Ca2+ gradient. Using KI and hypocrellin B (a photosensitive pigment obtained from a parasitic fungus, growing in Yunnan, China), the fluorescence quenching further indicated that the dynamic change of this tryptophan group, located at the protein-lipid interface, is a characteristic of transmembrane Ca2+ gradient-mediated conformational changes in SR · Ca2+-ATPase.Abbreviations SR sarcoplasmic reticulum - HB hypocrellin B - Trp tryptophan - DMSO dimethysulfoxide - Hepes N-2-hydroxyethyl piperazine-N-ethanesulfonic acad - SR(50005) SR vesicles with 1000-fold transmembrane Ca2+ gradient - SR(5050) SR vesicles without Ca2+ gradient - Ksv(app) apparent Stern-Volmer constant - Ksvi Stern-Volmer constant of component i for dynamic quenching  相似文献   

13.
The aggregation behavior of the carboxylic ionophore, lasalocid A, has been studied in egg phosphatidylcholine vesicles by monitoring the intrinsic fluorescence of lasalocid A. Self quenching of lasalocid A fluorescence in vesicles of egg phosphatidylcholine suggests aggregation of lasalocid A. When aggregated lasalocid A is treated with increasing concentrations of lipid, there is an increase in fluorescence due to gradual reduction of self quenching on lateral dilution. This confirms the presence of loosely held non-covalent aggregates of lasalocid A in the membrane. This result is relevant in elucidating the molecular mechanism of cation transport by lasalocid A across membranes.  相似文献   

14.
We investigated the permeabilities of lipid bilayers to the neurotoxin methylmercuric chloride (MMC). This mercurial is an efficient collisional quencher of the fluorescence of N-alkyl carbazole derivatives. Quenching of the fluorescence of β-(3-(9-carbazole)-propionyl-L--phosphatidylcholine (CPA-PC) in vesicles of dimyristoyl phosphatidylcholine and of dioleoyl phosphatidylcholine reveal rapid diffusion of MMC in the alkyl side chain regions of these bilayers. By a combination of (1) the lipid concentration dependence of the apparent quenching constants, (2) the solubility of MMC in concentrated lipid dispersions and (3) the 270 MHz proton magnetic resonance of methylmercury in the presence of lipid bilayers we conclude that the lipid-water partition coefficient of this mercurial is less than or equal to two. Using the fluorescence quenching and the partitioning data we estimate the diffusion coefficient of MMC in these bilayers to range from 0.13 to 0.31 × 10−5 cm2/sec, or 20–47% of its diffusion coefficient in ethanol. These data indicate that lipid bilayers do not pose a significant permeability barrier to the diffusional transport of MMC.  相似文献   

15.
Self- or concentration quenching of octadecylrhodamine B (C18-Rh) fluorescence increases linearly in egg phosphatidylcholine (PC) vesicles but exponentially in vesicles composed of egg PC:cholesterol, 1:1, as the probe concentration is raised to 10 mol%. Cholesterol-dependent enhancement of self-quenching also occurs when N-(lissamine-rhodamine-B-sulfonyl)dioleoylphosphatidylethanolamine is substituted for C18-Rh and resembles that in dipalmitoylphosphatidylcholine vesicles below, as opposed to above, the phase transition. These effects are not due to changes in dimer:monomer absorbance. Stern-Volmer plots indicate a dependence of quenching on nonfluorescent dimers both in the presence and absence of cholesterol. Decreases in fluorescence lifetimes with increasing probe concentration parallel decreases in residual fluorescence of C18-Rh with increasing probe concentration in PC and PC + cholesterol membranes, respectively. Decreases in the steady-state polarization of C18-Rh fluorescence as its concentration is raised to 10 mol% indicate energy transfer with emission between probe molecules in PC and to a lesser extent in PC + cholesterol membranes. The calculated R0 for 50% efficiency of energy transfer from excited state probe to monomer was 55-58 A and to dimer was 27 A. Since lateral diffusion of C18-Rh is probably too slow to permit collisional quenching during the lifetime of the probe, even if C18-Rh were concentrated in a separate phase, C18-Rh self-quenching appears to be due mainly to energy transfer without emission to nonfluorescent dimers.  相似文献   

16.
The quenching of core and core-shell CdSe quantum dots by TEMPO and 4-amino-TEMPO has been examined using steady state fluorescence spectroscopy. The efficiency of quenching is strongly dependent on the nanoparticle size, the binding properties of the nitroxide, and the presence or not of a protective shell, ZnS in our systems. The shell reduces the quenching efficiency significantly only in the case of binding nitroxides, such as 4-amino-TEMPO. Downward quenching plots revealing bimodal quenching characterize the Stern-Volmer plots obtained for 4-amino-TEMPO. The slope characteristic of the low concentrations regime depends strongly on the presence of a shell. For example, for particles with a 2.4 nm core, emitting at 525 nm the concentrations of 4-amino-TEMPO required to reduce the emission to one half are 0.65 microM and 0.08 mM for core and core-shell nanoparticles, respectively. Surprisingly, in the high concentration regime, a single Stern-Volmer slope of about 4000 M-1 seems to accommodate all systems. We speculate that this value is characteristic of the exchange of TOPO ligands by 4-amino-TEMPO.  相似文献   

17.
Quenching of tryptophan fluorescence by brominated phospholipid   总被引:7,自引:0,他引:7  
E J Bolen  P W Holloway 《Biochemistry》1990,29(41):9638-9643
Bromolipids [1-palmitoyl-2-(dibromostearoyl)phosphatidylcholine] with bromines at the 4,5-, 6,7-, 9,10-, 11,12-, and 15,16-positions were used to examine the fluorescence quenching of a synthetic, membrane-spanning peptide (Lys2-Gly-Leu8-Trp-Leu8-Lys-Ala-amide) incorporated into both small and large unilamellar vesicles. The peptide-lipid vesicles were analyzed to show that at least 75% of the peptide was in a transbilayer configuration, placing the single tryptophan in its predicted place in the center of the bilayer. Quenching profiles of the peptide in bromolipid showed maximal (90%) quenching by the 15,16-bromolipid, indicating that the bromolipids can accurately locate the position of a tryptophan in the bilayer. The quenching by the other bromolipids decreased with an r6 dependence and an apparent R0 of 9 A. In addition, indole in methanolic solution was subjected to quenching by a variety of mono- and dibrominated hydrocarbons. The quenching was analyzed, by using a modified Stern-Volmer equation, and found to be greatly dependent upon the number and positioning of the bromines. Monobromobutanes were found to have a quenching efficiency of only 7% while dibromobutanes, with bromines on adjacent carbon atoms, had efficiencies of over 80%. In addition, the dibromobutanes exhibited significant "static" quenching whereas the monobrominated butanes did not. These data suggest that the bromolipids are more appropriately defined as short-range quenchers rather than strictly contact quenchers.  相似文献   

18.
Anilinonaphtyl labeled spectrin exhibits a fluorescence emission spectrum characteristic of a highly hydrophobic environment. Quenching of the fluorescence intensity by nitroxide analogs of fatty acids of affinity 10(4) M-1 reveals that the sites of interaction of fatty acids lie very close to the anilinonaphtyl groups. Similar experiments performed with a nitroxide analog of phosphatidylserine yield a 30% quenching of fluorescence while the same phosphatidylcholine analog has essentially no effect. The changes in the fluorescence emission spectrum exhibited in the presence of sonicated phosphatidylserine vesicles further outline the specificity of interaction towards phosphatidylserine, with one spectrin binding site per about 750 exposed phospholipids. Moreover, they suggest a penetration of the anilinonaphtyl group into the lipid bilayer.  相似文献   

19.
Lipophilin, a hydrophobic protein purified from the proteolipid of normal human brain myelin, was recombined with phosphatidylcholine by solubilization of the lipid and protein in 2-chloro-ethanol followed by dialysis against buffer. This method resulted in homogeneous incorporation of the protein into lipid vesicles as judged by sedimentation on a sucrose gradient and freeze fracture electron microscopy. The lipid-protein vesicles were single layered, 1000–2000 Å in diameter and the freeze fracture faces contained intramembrane particles. The effect of lipophilin on the organization of the lipid was studied by use of spin label probes. Two distinct components were present in the spectrum of fatty acid spin labels in the lipid-protein vesicles. One was immobilized presumably due to the presence of boundary lipid around the protein and the second component was indicative of aniostropic motion similar to the spectrum in phosphatidylcholine vesicles and probably due to a lamellar phase but with a slightly greater order parameter. Lipophilin was found to increase the order parameter linearly with increasing concentration of protein incorporated into the vesicles. However, the phase transition temperature as measured from the 2,2,6,6-tetramethyl piperidine-1-oxyl (TEMPO) solubility parameter was unchanged.  相似文献   

20.
Fluorescence quenching techniques have been used extensively in recent years to examine reaction rates and the compartmentalization of components in lipid micelles and membranes. Steady-state fluorescence methods are frequently employed in such studies but the interpretation of the resulting Stern-Volmer plots is often hampered by uncertainties regarding the mode of association of the quencher with the lipid structure and the nature of the quenching mechanism. This paper presents a method for simulating steady-state Stern-Volmer plots in two phase systems, and shows how the forms of such plots are influenced by the type of association of the quencher with the membrane or micelle (partition and/or binding) and by the type of quenching mechanism (dynamic and/or static). Comparisons of simulated plots with experimental data must take into account the possible combinations of quencher association(s) and quenching mechanism(s). The methods presented are applicable to synthetic and natural membranes and provide a basis for comparing the quenching of fluorescent molecules in biological membranes of differing composition.  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