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.
A reference method for the deconvolution of polarized fluorescence decay data is described. Fluorescence lifetime determinations for p-terphenyl, p-bis[2-(5-phenyloxazolyl)]benzene and N-acetyltryptophanamide (AcTrpNH2) show that with this method more reliable fits of the decays can be made than with the scatterer method, which is most frequently used. Analysis of the AcTrpNH2 decay with p-terphenyl as the reference compound yields an excellent fit with lifetimes of 2.985 ns for AcTrpNH2 and 1.099 ns for p-terphenyl (20 degrees C), whereas the AcTrpNH2 decay cannot be satisfactorily fitted when the scatterer method is used. The frequency of the detected photons is varied to determine the conditions where pulse pile-up starts to affect the measured decays. At detection frequencies of 5 kHz and 15 kHz, which corresponds to 1.7% and 5% respectively of the rate of the excitation photons no effects are found. Decays measured at 30 kHz (10%) are distorted, indicating that pile-up effects play a role at this frequency. The fluorescence and fluorescence anisotropy decays of the tryptophan residues in the proteins human serum albumin, horse liver alcohol dehydrogenase and lysozyme have been reanalysed with the reference method. The single tryptophan residue of the albumin is shown to be characterized by a triple-exponential fluorescence decay. The anisotropy decay of albumin was found to be mono-exponential with a rotational correlation time of 26 ns (20 degrees C). The alcohol dehydrogenase has two different tryptophan residues to which single lifetimes are assigned. It is found that the rotational correlation time for the dehydrogenase changes with excitation wavelength (33 ns for lambda ex = 295 nm and 36 ns for lambda ex = 300 nm at 20 degrees C), indicating a nonspherical protein molecule. Lysozyme has six tryptophan residues, which give rise to a triple-exponential fluorescence decay. A single-exponential decay with a rotational correlation time of 3.8 ns is found for the anisotropy. This correlation time is significantly shorter than that arising from the overall rotation and probably originates from intramolecular, segmental motion.  相似文献   

2.
The measurement of fluorescence lifetimes is known to be hindered by the wavelenght-dependent and photocathode area-dependent time response of photomultiplier tubes. A simple and direct method is described to minimize the effects in photomultiplier tubes for phase-modulation fluorometry. Reference fluorophores of known lifetime were used in place of the usual scattering reference. The emission wavelenghts of the reference and sample were matched by either filters or a monochromator, and the use of a fluorophore rather than a scatter decreases the differences in spatial distribution of light emanating from the reference and sample. Thus photomultiplier tube artifacts are minimized. Five reference fluorophores were selected on the basis of availability, ease of solution preparation, and constancy of lifetime with temperature and emission wavelenght. These compounds are p-terphenyl, PPO, PPD, POPOP and dimethyl POPOP. These compounds are dissolved in ethanol to give standard solutions that can be used over the temperature range from ?55 to +55°C. Purging with inert gas is not necessary. The measured phase and modulation of the reference solution is used, in conjunction with the known reference, lifetime, to calculate the actual phase and modulation of the exictation beam. The use of standard fluorophores does not require separate experiments to quantify photomultiplier effects, and does not increase the time required for the measurement of fluorescence lifetimes. Examples are presented which demonstrate the elimination of artifactual photomultiplier effects in measurements of the lifetimes of DADH (0.4 ns) and indole solutions quenched by iodide. In addition, the use of these reference solutions increases the accuracy of fluorescence lifetime measurements ranging ranging to 30 ns. We judge this method to provide more reliable lifetime measurements by the phase and modulation method. The test solutions and procedures we describe may be used by other laboratories to evaluate the performance of their phase fluorometers.  相似文献   

3.
We have used one- (OPE) and two-photon (TPE) excitation with time-correlated single-photon counting techniques to determine time-resolved fluorescence intensity and anisotropy decays of the wild-type Green Fluorescent Protein (GFP) and two red-shifted mutants, S65T-GFP and RSGFP. WT-GFP and S65T-GFP exhibited a predominant approximately 3 ns monoexponential fluorescence decay, whereas for RSGFP the main lifetimes were approximately 1.1 ns (main component) and approximately 3.3 ns. The anisotropy decay of WT-GFP and S65T-GFP was also monoexponential (global rotational correlation time of 16 +/- 1 ns). The approximately 1.1 ns lifetime of RSGFP was associated with a faster rotational depolarization, evaluated as an additional approximately 13 ns component. This feature we attribute tentatively to a greater rotational freedom of the anionic chromophore. With OPE, the initial anisotropy was close to the theoretical limit of 0.4; with TPE it was higher, approaching the TPE theoretical limit of 0.57 for the colinear case. The measured power dependence of the fluorescence signals provided direct evidence for TPE. The general independence of fluorescence decay times, rotation correlation times, and steady-state emission spectra on the excitation mode indicates that the fluorescence originated from the same distinct excited singlet states (A*, I*, B*). However, we observed a relative enhancement of blue fluorescence peaked at approximately 440 nm for TPE compared to OPE, indicating different relative excitation efficiencies. We infer that the two lifetimes of RSGFP represent the deactivation of two substates of the deprotonated intermediate (I*), distinguished by their origin (i.e., from A* or B*) and by nonradiative decay rates reflecting different internal environments of the excited-state chromophore.  相似文献   

4.
In wavelength-resolved steady state spectra we observe three different kinds of emission from histone H1, a class A protein with only a single tyrosine residue. Unfolded H1 emissions that peak at approximately 300 and 340 nm can both be excited maximally at approximately 280 nm. Another, peaking much further to the red at approximately 400 nm, can be excited maximally at approximately 320 nm. The 300-nm fluorescence can be resolved by lifetime measurements into three components with decay times of approximately 1, 2, and 4 ns. On sodium-chloride-induced refolding of H1, simplification of the emission properties occurs. The 340 and 400-nm components disappear while the two shorter lifetime components of the 300-nm band diminish in amplitude and are replaced by the 4-ns decay. We believe that the 340-nm emission is tyrosinate fluorescence resulting from excited-state proton transfer. The origin of the 400-nm emission remains uncertain. We assign the 1 and 2-ns components of the 300-nm emission to two states of tyrosine in denatured H1 and the 4-ns decay to fluorescence of the single tyrosine residue in the globular region of refolded H1. Our results support the contention that salt induced folding of H1 is a cooperative two state process, and permit us to better understand the previously reported increases in fluorescence intensity and anisotropy on salt-induced folding.  相似文献   

5.
It is shown that all-trans-retinal under model conditions of its excessive accumulation in photoreceptor membranes interacts with amino groups of rhodopsin and lipids, forming at least three distinct fluorophores with fluorescence quantum yield 20–40 times higher than that of free all-trans-retinal. These retinal derivatives are likely precursors of photo- and cytotoxic fluorophores of lipofuscin and in particular of A2E. Spectral characteristics of fluorophores have been described. Picosecond time-resolved laser fluorescence spectroscopy was used to study kinetics of fluorescence decay of both free and bound all-trans-retinal; fluorophores were determined and their lifetimes have been measured. Based on calculations it is shown that the decay kinetics of all-trans-retinal derivatives consists of three components with lifetimes equal to 48, 208, and 900 ps; kinetics of free all-trans-retinal is monoexponential with lifetime of 31 ps. The chemical nature of fluorophores with the lifetimes obtained is discussed.  相似文献   

6.
We have compared the performance of two Troponin-C-based calcium FRET sensors using fluorescence lifetime read-outs. The first sensor, TN-L15, consists of a Troponin-C fragment inserted between CFP and Citrine while the second sensor, called mTFP-TnC-Cit, was realized by replacing CFP in TN-L15 with monomeric Teal Fluorescent Protein (mTFP1). Using cytosol preparations of transiently transfected mammalian cells, we have measured the fluorescence decay profiles of these sensors at controlled concentrations of calcium using time-correlated single photon counting. These data were fitted to discrete exponential decay models using global analysis to determine the FRET efficiency, fraction of donor molecules undergoing FRET and calcium affinity of these sensors. We have also studied the decay profiles of the donor fluorescent proteins alone and determined the sensitivity of the donor lifetime to temperature and emission wavelength. Live-cell fluorescence lifetime imaging (FLIM) of HEK293T cells expressing each of these sensors was also undertaken. We confirmed that donor fluorescence of mTFP-TnC-Cit fits well to a two-component decay model, while the TN-L15 lifetime data was best fitted to a constrained four-component model, which was supported by phasor analysis of the measured lifetime data. If the constrained global fitting is employed, the TN-L15 sensor can provide a larger dynamic range of lifetime readout than the mTFP-TnC-Cit sensor but the CFP donor is significantly more sensitive to changes in temperature and emission wavelength compared to mTFP and, while the mTFP-TnC-Cit solution phase data broadly agreed with measurements in live cells, this was not the case for the TN-L15 sensor. Our titration experiment also indicates that a similar precision in determination of calcium concentration can be achieved with both FRET biosensors when fitting a single exponential donor fluorescence decay model to the fluorescence decay profiles. We therefore suggest that mTFP-based probes are more suitable for FLIM experiments than CFP-based probes.  相似文献   

7.
We describe a new method for the analysis of phase-sensitive fluorescence emission spectra. This method permits the resolution of three-component mixtures using spectra measured at a single modulation frequency. Phase-sensitive spectra are recorded using one modulation frequency, at a number of arbitrary detector phase angles. It is not necessary to suppress any one component. The spectra are then used to estimate the component lifetimes and steady-state fractional intensities using a nonlinear least-squares analysis procedure. The only requirement for the analysis is the knowledge of the steady-state spectra of the individual components. This procedure allowed the resolution of a two-component mixture of 9-methylanthracene (4.5 ns) and 9,10-diphenylanthracene (5.9 ns). It should be noted that resolution of two lifetimes which differ by only 30% is a difficult task. Additionally, we resolved a three-component mixture with lifetimes that differed fourfold: p-bis[2-(5-phenyloxazolyl)]benzene (1.3 ns), 9-methylanthracene (4.5 ns), and 9,10-diphenylanthracene (5.9 ns). Conveniently, the technique utilizes a commercially available fixed-frequency phase fluorometer.  相似文献   

8.
The binding of cyclosporine to human peripheral blood lymphocytes (PBLs) was studied by measuring the fluorescence emission spectrum and lifetime of the fluorescent and immunosuppressive cyclosporine derivative dansyl-cyclosporine (DCs). The emission maximum and fluorescence lifetime of DCs were characterized in several solvents. The fluorescence emission maximum and lifetime of DCs increased at a high dielectric constant. The fluorescence lifetime decay curve of DCs was a monoexponential function in all solvents tested. Fluorescence micrographs of lipid vesicles and erythrocytes labeled with DCs exhibit uniform staining patterns, whereas PBLs show heterogeneous DCs labeling. DCs exhibits a relatively low emission maximum (490 nm) in erythrocyte membranes. Such an emission maximum is characteristic of a hydrophobic environment. DCs in PBLs also has a low emission maximum (484 nm). The lifetime of DCs in PBLs required two exponential terms to properly fit the lifetime decay curve and could not be attributed to light scattering. One short component (4.7 +/- 1.0 ns) and a second long component (18.5 +/- 1.0 ns) were resolved from the DCs fluorescence decay curves. Time-resolved anisotropy of DCs in PBLs revealed that the labeled drug was present in an anisotropic environment, consistent with at least some DCs being bound to a membrane. These fluorescence studies suggest that DCs interacts with multiple and/or heterogeneous sites in peripheral blood lymphocytes.  相似文献   

9.
The fluorescence lifetime of the single tryptophan in whiting parvalbumin has been measured by time-correlated single-photon counting. In the presence of saturating calcium, greater than 2 mol/mol of protein, the decay of fluorescence is accurately single exponential with a lifetime of 4.6 ns (0.1 M KCl, 20 mM borate, 1 mM dithiothreitol, 20 degrees C, pH 9). Upon complete removal of calcium from parvalbumin with ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid the emission decay becomes biphasic, and a second more rapid decay process with a lifetime of 1.3 ns comprising approximately 18% of the fluorescence emission at 350 nm is observed. The fluorescence emission of the calcium-saturated form is not measurably quenched by iodide. In contrast, upon complete removal of calcium, the fluorescence is completely quenchable as shown by extrapolation of the data to infinite iodide concentration. These results indicate that there is a large increase in the accessibility of the tryptophan residue in the protein to solvent upon removal of calcium. Stern-Volmer plots of the quenching data are nonlinear and indicate that there is more than one quenchable conformation of the calcium-free protein. The lifetime and quenching results are consistent with the presence of significant concentrations of only two stoichiometric species, apoparvalbumin and parvalbumin--Ca2, at partial occupancy of the calcium binding sites.  相似文献   

10.
The biodistribution of two near-infrared fluorescent agents was assessed in vivo by time-resolved diffuse optical imaging. Bacteriochlorophyll a (BC) and cypate-glysine-arginine-aspartic acid-serine-proline-lysine-OH (Cyp-GRD) were administered separately or combined to mice with subcutaneous xenografts of human breast adenocarcinoma and slow-release estradiol pellets for improved tumor growth. The same excitation (780 nm) and emission (830 nm) wavelengths were used to image the distinct fluorescence lifetime distribution of the fluorescent molecular probes in the mouse cancer model. Fluorescence intensity and lifetime maps were reconstructed after raster-scanning whole-body regions of interest by time-correlated single-photon counting. Each captured temporal point-spread function (TPSF) was deconvolved using both a single and a multiexponental decay model to best determine the measured fluorescence lifetimes. The relative signal from each fluorophore was estimated for any region of interest included in the scanned area. Deconvolution of the individual TPSFs from whole-body fluorescence intensity scans provided corresponding lifetime images for comparing individual component biodistribution. In vivo fluorescence lifetimes were determined to be 0.8 ns (Cyp-GRD) and 2 ns (BC). This study demonstrates that the relative biodistribution of individual fluorophores with similar spectral characteristics can be compartmentalized by using the time-domain fluorescence lifetime gating method.  相似文献   

11.
We describe the construction and performance of a variable-frequency phase-modulation fluorometer. This instrument, which provides modulation frequencies from 1 to 200 MHz, was constructed using commercially available components. To facilitate the introduction of these instruments into other laboratories we describe in detail the chosen components and the principles of operation. The present light source is a continuous-wave helium-cadmium laser, which provides convenient excitation wavelengths of 325 and 442 nm. Modulation of the incident light is provided by one of several electro-optic modulators. The extent of modulation ranges from 1.0 to 0.2 as the frequency increases from 1 to 200 MHz. Phase angles and demodulation factors are measured using the cross-correlation method. The closely spaced frequencies are provided by two direct frequency synthesizers. The phase and modulation measurements are accurate to 0.2 degrees and 0.002, respectively, from 1 to 200 MHz. This accuracy allows considerable resolution of complex decay laws. The usefulness of frequency-domain fluorometry for the resolution of multiexponential decays is illustrated by the analysis of several difficult mixtures. As examples, we resolved a two-component mixture of anthracene (4.1 ns) and 9,10-diphenylanthracene (6.3 ns), and confirmed that the intensity decay of NADH in aqueous buffer is at least a double exponential (0.2 and 0.86 ns). We also resolved an especially difficult mixture of anthracene (4.1 ns) and 9-methylanthracene (4.5 ns), and a three-component mixture with decay times of 1.3, 4.1 and 7.7 ns. Frequency-domain fluorometers appear to be particularly useful for determination of complex decays of fluorescence anisotropy. This capability is illustrated by the determination of rotational correlation times as short as 47 ps for p-bis[2-(5-phenyloxazolyl)]benzene (POPOP) in hexane at 40 degrees C, and by the resolution of the two correlation times of anisotropic rotators such as perylene and 9-aminoacridine. Resolution of two anisotropy decay times for 9-aminoacridine is a difficult test because these correlation times differ by less than 2-fold. The resolution of multiexponential decays of intensity and anisotropy possible with this instrument is at least equivalent to that obtained using state-of-the-art time-resolved instruments based on mode-locked laser sources. The ease and rapidity of frequency-domain measurements, the relative simplicity of the equipment, the accuracy of the measurements and the lack of significant systematic errors indicate that frequency-domain fluorometry will be widely useful in chemical and biochemical research.  相似文献   

12.
13.
The fluorescence decay of apoazurin derived from Pseudomonas aeruginosa is monoexponential. By this criterion the population of molecules of apoazurin is homogeneous. The emission anisotropy factor and the absorption anisotropy factor at the red edge of the absorption band assume similar values, showing that the tryptophan residue in apoazurin has the same asymmetric environment both in the ground and excited states. This finding suggests tight packing of the protein at the tryptophan environment. Native azurin does not decay monoexponentially. Moreover, comparison between the quantum yield calculated from the decay kinetics and the one measured directly shows that the majority of the azurin molecules are not fluorescent. There is thus variability in the structure of azurin molecules with an equilibration time that is longer than the fluorescence lifetime. Different asymmetric environment was found for the tryptophan residue in oxidized and reduced holoprotein and in apoazurin, as studied by the circular polarization of the fluorescence. D(2)O increases the fluorescence lifetime of apoazurin by 6 percent, compared to the lifetime in H(2)O solution; therefore water molecules may have access to the tryptophan residue, though the latter is situated in a hydrophobic environment.  相似文献   

14.
D Amir  E Haas 《Biochemistry》1987,26(8):2162-2175
A series of four bovine pancreatic trypsin inhibitor (BPTI) derivatives, site specifically labeled by (2-methoxy-1-naphthyl)methyl (MNA) at the N-terminal amino group and by [7-(dimethylamino)-coumarin-4-yl]acetyl (DA-coum) at one of the four epsilon-amino groups, was prepared. The four derivatives, N alpha-MNA-Arg1-N epsilon-DA-coum-Lysn-BPTI [(1-n)BPTI] (n = 15, 26, 41, and 46), were purified by affinity chromatography and high-performance liquid chromatography (HPLC). The homogeneity of each derivative and its site of labeling were characterized by HPLC tryptic peptide mapping. Nonradiative energy transfer from MNA (donor) to DA-coum (acceptor) was measured by monitoring donor emission and acceptor excitation spectra. Transfer efficiencies between 45% and 85% were observed. The fluorescence decay of MNA in MNA-BPTI, a derivative labeled by a donor without an acceptor, is monoexponential, with a lifetime of 6.8 +/- 0.15 ns. The decay kinetics of MNA fluorescence measured for derivatives labeled both by donor and acceptor showed a small deviation from monoexponential decay with shorter average lifetimes. Analysis of the experimental decay curves yielded the detailed intramolecular distance distribution functions for each pair of labeled sites. The averages of the calculated distance distribution functions are close to the values expected from the known structure of BPTI in the crystalline state. The derivatives thus obtained are suitable for investigation of conformational transitions of the labeled protein and for monitoring localized changes such as those involved in the folding or unfolding transitions.  相似文献   

15.
Hydrodynamic properties as well as structural dynamics of proteins can be investigated by the well-established experimental method of fluorescence anisotropy decay. Successful use of this method depends on determination of the correct kinetic model, the extent of cross-correlation between parameters in the fitting function, and differences between the timescales of the depolarizing motions and the fluorophore's fluorescence lifetime. We have tested the utility of an independently measured steady-state anisotropy value as a constraint during data analysis to reduce parameter cross correlation and to increase the timescales over which anisotropy decay parameters can be recovered accurately for two calcium-binding proteins. Mutant rat F102W parvalbumin was used as a model system because its single tryptophan residue exhibits monoexponential fluorescence intensity and anisotropy decay kinetics. Cod parvalbumin, a protein with a single tryptophan residue that exhibits multiexponential fluorescence decay kinetics, was also examined as a more complex model. Anisotropy decays were measured for both proteins as a function of solution viscosity to vary hydrodynamic parameters. The use of the steady-state anisotropy as a constraint significantly improved the precision and accuracy of recovered parameters for both proteins, particularly for viscosities at which the protein's rotational correlation time was much longer than the fluorescence lifetime. Thus, basic hydrodynamic properties of larger biomolecules can now be determined with more precision and accuracy by fluorescence anisotropy decay.  相似文献   

16.
Monitoring fluorescence properties of endogenous fluorophores such as nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FAD) in normal and cancerous cells provide substantial information noninvasively on biochemical and biophysical aspects of metabolic dysfunction of cancerous cells. Time‐resolved spectral profiles and fluorescence lifetime images of NADH and FAD were obtained in human lung nonsmall carcinomas (H661 and A549) and normal lung cells (MRC‐5). Both fluorophores show the fast and slowly decaying emission components upon pulsed excitation, and fluorescence spectra of NADH and FAD show blue‐ and red‐shifts, respectively, during their decay. All identified lifetime components of NADH and FAD were found to be shorter in cancerous cells than in normal cells, no matter how they were measured under different extra‐cellular conditions (cells suspended in cuvette and cells attached on glass substrate), indicating that the changes in metabolism likely altered the subcellular milieu and potentially also affected the interaction of NADH and FAD with enzymes to which these cofactors were bound. The intensity ratio of NADH and FAD of cancerous cells was also shown to be larger than that of normal cells.  相似文献   

17.
The interaction of several 3,6-diaminoacridines with DNAs of various base composition has been studied by steady-state and transient fluorescence measurements. The acridine dyes employed are of the following two classes: class I - proflavine, acriflavine and 10-benzyl proflavine; class II - acridine yellow, 10-methyl acridine yellow and benzoflavine. It is found that the fluorescence decay kinetics follows a single-exponential decay law for free dye and the poly[d(A-T)]-dye complex, while that of the dye bound to DNA obeys a two-exponential decay law. The long lifetime (tau 1) for each complex is almost the same as the lifetime for the poly[d(A-T)]-dye complex, and the amplitude alpha 1 decreases with increasing GC content of DNA. The fluorescence quantum yields (phi F) of dye upon binding to DNA decrease with increasing GC content; the phi F values for class I are nearly zero when bound to poly(dG) X poly(dC), but those for class II are not zero. This is in harmony with the finding that GMP almost completely quenches the fluorescence for class I, whereas a weak fluorescence arises from the GMP-dye complex for class II. The fluorescence spectra of the DNA-dye complexes gradually shift toward longer wavelengths with increasing GC content. In this connection, the fluorescence decay parameters show a dependence on the emission wavelength; alpha 1 decreases with an increase in the emission wavelength. In view of these results, it is proposed that the decay behavior of the DNA-dye complexes has its origin in the heterogeneity of the emitting sites; the long lifetime tau 1 results from the dye bound to AT-AT sites, while the short lifetime tau 2 is attributable to the dye bound in the vicinity of GC pairs. Since GC pairs almost completely quench the fluorescence for class I, partly intercalated or externally bound dye molecules may play an important role in the component tau 2.  相似文献   

18.
A microscope based time-correlated single photon counting instrument has been constructed to measure fluorescence intensity and emission anisotropy decays from fluorophores in single cells on a nanosecond time scale. The sample is excited and the emission collected using epi-illumination optics with frequency-doubled pulses from the cavity-dumped output of a synchronously pumped dye laser serving as an excitation source. Collection of decays from a single cell is possible due to the presence of an iris in the emission path that can be reduced to less than the diameter of a single cell. Using the instrument the decay of 60 nM 1,6-diphenyl-1,3,5-hexatriene was measured, demonstrating that adequate data for lifetime analysis can be recorded from fewer 10(3) molecules of the fluorophore in an illuminated volume of 23 fl. In addition, the intensity and anisotropy decays of fura-2 in single adherent cells and in suspensions of fura-2 loaded cells in suspension, although the relative amplitudes and decay constants vary somewhat from cell to cell. The results indicate that a significant but variable fraction of fura-2 is bound to relatively immobile macromolecular components in these cells.  相似文献   

19.
Physical properties of the fluorescent sterol probe dehydroergosterol   总被引:3,自引:0,他引:3  
Spectroscopic studies were performed on the fluorescent sterol probes ergosta-5,7,9(11),22-tetraen-3 beta-ol (dehydroergosterol) and cholesta-5,7,9(11)-trien-3 beta-ol (cholestatrienol). In most isotropic solvents, these molecules exhibited a single lifetime near 300 ps. Fluorescence lifetimes in 2-propanol were independent of emission wavelength and independent of excitation wavelength. Excited state behavior of these probes appears relatively simple. In isotropic solvents, dehydroergosterol fluorescence emission underwent at most a small Stokes shift as solvent polarity was modified. Time-resolved anisotropy decays indicated that dehydroergosterol decay was monoexponential, with rotational correlation times dependent on solvent viscosity. When incorporated into L-alpha-dimyristoylphosphatidylcholine liposomes at a concentration of 0.9 mol%, dehydroergosterol fluorescence lifetime decreased at the phase transition of this phospholipid indicating that the sterol probe was detecting physical changes of the bulk phospholipids. Furthermore, total fluorescence decays and anisotropy decays were sensitive to the environment of the sterol. Dehydroergosterol and cholestatrienol are thus useful probes for monitoring sterol behavior in biological systems.  相似文献   

20.
Förster resonant energy transfer (FRET) measurements are widely used to obtain information about molecular interactions and conformations through the dependence of FRET efficiency on the proximity of donor and acceptor fluorophores. Fluorescence lifetime measurements can provide quantitative analysis of FRET efficiency and interacting population fraction. Many FRET experiments exploit the highly specific labelling of genetically expressed fluorescent proteins, applicable in live cells and organisms. Unfortunately, the typical assumption of fast randomization of fluorophore orientations in the analysis of fluorescence lifetime‐based FRET readouts is not valid for fluorescent proteins due to their slow rotational mobility compared to their upper state lifetime. Here, previous analysis of effectively static isotropic distributions of fluorophore dipoles on FRET measurements is incorporated into new software for fitting donor emission decay profiles. Calculated FRET parameters, including molar population fractions, are compared for the analysis of simulated and experimental FRET data under the assumption of static and dynamic fluorophores and the intermediate regimes between fully dynamic and static fluorophores, and mixtures within FRET pairs, is explored. Finally, a method to correct the artefact resulting from fitting the emission from static FRET pairs with isotropic angular distributions to the (incorrect) typically assumed dynamic FRET decay model is presented.   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