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.
Fluorescence recovery after photobleaching with an unmodified confocal laser scanning microscope (confocal FRAP) was used to determine the diffusion properties of network forming biological macromolecules such as aggrecan. The technique was validated using fluorescein isothiocyanate (FITC)-labeled dextrans and proteins (molecular mass 4-2000 kDa) at 25 degrees C and with fluorescent microspheres (207 nm diameter) over a temperature range of 5-50 degrees C. Lateral diffusion coefficients (D) were independent of the focus position, and the degree and extent of bleach. The free diffusion coefficient (Do) of FITC-aggrecan determined by confocal FRAP was 4.25 +/- 0.6 x 10(-8) cm2 s-1, which is compatible with dynamic laser light scattering measurements. It appeared to be independent of concentration below 2.0 mg/ml, but at higher concentrations (2-20 mg/ml) the self-diffusion coefficient followed the function D = Do(e)(-Bc). The concentration at which the self-diffusion coefficient began to fall corresponded to the concentration predicted for domain overlap. Multimolecular aggregates of aggrecan ( approximately 30 monomers) had a much lower free diffusion coefficient (Do = 6.6 +/- 1.0 x 10(-9) cm2 s-1) but showed a decrease in mobility with concentration of a form similar to that of the monomer. The method provides a technique for investigating the macromolecular organization in glycan-rich networks at concentrations close to those found physiologically.  相似文献   

2.
We studied intracellular binding and possible compartmentalization of the fluorescent Ca2+ indicators, indo-1 and fura-2, in single mammalian cardiac ventricular cells that had been loaded with indo-1 and fura-2 by exposure to the acetoxymethylester form of the indicators (indo-1/AM and fura-2/AM). Techniques similar to those used in experiments on fluorescence recovery after photobleaching (FRAP) were used. It was assumed that reversible binding in myoplasm would be evident as slowed recovery of fluorescence after photobleaching, and that irreversible binding of the indicators to immobile myoplasmic sites (or "compartmentalization" in organelles) would be evident as incomplete recovery. Through the use of a mask, one half of a cell was exposed to high-intensity ultraviolet (UV) light to bleach the indo-1 or fura-2 in only that part of the cell. Upon removal of the mask and termination of the high-intensity UV illumination, fluorescence recovered in the bleached half of the cell, indicating diffusion of indo-1 and fura-2. Mathematical modeling of the diffusional redistribution of the indicators indicated that in these cells the apparent diffusion coefficient for indo-1 is 1.57 x 10(-7) cm2 s-1 (SD 0.48 x 10(-7) cm2 s-1; n = 5 cells, 21 degrees C), and for fura-2 is 3.19 x 10(-7) cm2 s-1 (SD 1.85 x 10(-7) cm2 s-1; n = 6 cells, 21 degrees C). These values are approximately 6 and 3, respectively, times smaller than those expected for free diffusion in the myoplasm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Color variants of green fluorescent protein (GFP) are increasingly used for multicolor imaging, fluorescence resonance energy transfer (FRET), and fluorescence recovery after photobleaching (FRAP). Here we show that experimental settings commonly used in these imaging experiments may induce an as yet uncharacterized reversible photobleaching of fluorescent proteins, which is more pronounced at acidic pH. Whereas the reversible photobleaching spectrum of eCFP corresponds to its absorption spectrum, reversible photobleaching spectra of yellow variants resemble absorption spectra of their protonated states. Fluorescence intensities recover spontaneously with time constants of 25-58 s. The recovery of eCFP can be further accelerated by illumination. The resulting steady-state fluorescence reflects a variable equilibrium between reversible photobleaching, spontaneous recovery, and light-induced recovery. These processes can cause significant artifacts in commonly applied imaging techniques, photobleach-based FRET determinations, and FRAP assays.  相似文献   

4.
Fluorescence recovery after photobleaching (FRAP) experiments to measure the mobility of cell surface components require a brief, but intense, pulse of light to photobleach the fluorescence in a restricted area of the cell. We studied possible photodamage to the cell surface during the photobleaching step using light and scanning electron microscopy (SEM) and various FRAP measurements themselves. The cell membrane was left impermeable to trypan blue after photobleaching. SEM studies show that the morphology of the cell surface is not altered by photobleaching. Cells can be repeatedly photobleached and/or photobleached using longer bleach times and greater intensities without systematically altering FRAP kinetics. Singlet oxygen quenchers or free radical traps designed to inhibit putative photoreagents produced during photobleaching do not markedly affect the results. Fluorescein and rhodamine labels give similar results. All of these results, obtained with several different monolayer cultures, suggest that photodamage induced during photobleaching is not a serious artefact in the cellular FRAP results obtained to date.  相似文献   

5.
Total internal reflection-fluorescence recovery after photobleaching (TIR-FRAP) was applied to measure solute translational diffusion in the aqueous phase of membrane-adjacent cytoplasm. TIR fluorescence excitation in aqueous solutions and fluorescently labeled cells was produced by laser illumination at a subcritical angle utilizing a quartz prism; microsecond-resolution FRAP was accomplished by acousto-optic modulators and electronic photomultiplier gating. A mathematical model was developed to determine solute diffusion coefficient from the time course of photobleaching recovery, bleach time, bleach intensity, and evanescent field penetration depth; the model included irreversible and reversible photobleaching processes, with triplet state diffusion. The validity and accuracy of TIR-FRAP measurements were first examined in aqueous fluorophore solutions. Diffusion coefficients for fluorescein isothiocyanate-dextrans (10-2000 kDa) determined by TIR-FRAP (recovery t1/2 0.5-2.2 ms) agreed with values measured by conventional spot photobleaching. Model predictions for the dependence of recovery curve shape on solution viscosity, bleach time, and bleach depth were validated experimentally using aqueous fluorescein solutions. To study solute diffusion in cytosol, MDCK epithelial cells were fluorescently labeled with the small solute 2',7'-bis-2-carboxyethyl-5-carboxyfluorescein-acetoxymethyl-ester (BCECF). A reversible photobleaching process (t1/2 approximately 0.5 ms) was identified that involved triplet-state relaxation and could be eliminated by triplet-state quenching with 100% oxygen. TIR-FRAP t1/2 values for irreversible BCECF bleaching, representing BCECF translational diffusion in the evanescent field, were in the range 2.2-4.8 ms (0.2-1 ms bleach times), yielding a BCECF diffusion coefficient 6-10-fold less than that in water. These results establish the theory and the first experimental application of TIR-FRAP to measure aqueous-phase solute diffusion, and indicate slowed translational diffusion of a small solute in membrane-adjacent cytosol.  相似文献   

6.
Frog rod outer segments were labeled with the sulfhydryl-reactive label iodoacetamido tetramethylrhodamine. The bulk of the label reacted with the major disk membrane protein, rhodopsin. Fluorescence photobleaching and recovery (FPR) experiments on labeled rods showed that the labeled proteins diffused rapidly in the disk membranes. In these FPR experiments we observed both the recovery of fluorescence in the bleached spot and the loss of fluorescence from nearby, unbleached regions of the photoreceptor. These and previous experiments show that the redistribution of the fluorescent labeled proteins after bleaching was due to diffusion. The diffusion constant, D, was (3.0 +/- 10(-9) cm2 s-1 if estimated from the rate of recovery of fluorescence in the bleached spot, and (5.3 +/- 2.4) x 10(-9) cm2 s-1 if estimated from the rate of depletion of fluorescence from nearby regions. The temperature coefficient, Q10, for diffusion was 1.7 +/- 0.5 over the range 10 degrees--29 degrees C. These values obtained by FPR are in good agreement with those previously obtained by photobleaching rhodopsin in fresh, unlabeled rods. This agreement indicates that the labeling and bleaching procedures required by the FPR method did not significantly alter the diffusion rate of rhodopsin. Moreover, the magnitude of the diffusion constant for rhodopsin is that to be expected for an object of its diameter diffusing in a bilayer with the viscosity of the disk membrane. In contrast to the case of rhodopsin, FPR methods applied to other membrane proteins have yielded much smaller diffusion constants. The present results help indicate that these smaller diffusion constants are not artifacts of the method but may instead be due to interactions the diffusing proteins have with other components of the membrane in addition to the viscous drag imposed by the lipid bilayer.  相似文献   

7.
Fluorescence photobleaching recovery with total internal reflection illumination (TIR-FPR) has been used to measure the dissociation kinetics of a fluorescein-labeled anti-dinitrophenyl monoclonal Fab specifically bound to supported monolayers composed of a mixture of dipalmitoylphosphatidylcholine and dinitrophenyl-conjugated dipalmitoylphosphatidylethanolamine. The fluorescence recovery curves were not monoexponential; when analyzed as a sum of two exponentials, the rates and fractional recoveries were approximately 1 s-1 (approximately 50%) and approximately 0.1 s-1 (approximately 30%). The data did not change as a function of the Fab solution concentration, indicating that the fluorescence recovery curves were not influenced by the rate of diffusion in bulk solution. Also, the recovery curves were independent of the size of the illuminated area, indicating that surface diffusion did not significantly contribute to the rate and shape of the fluorescence recovery. The measured off rates and apparent association constant (1.6 x 10(5) M-1) were analyzed with the theoretical formalism for a proposed mechanism that accounts for the nonmonoexponential kinetics.  相似文献   

8.
Although fluorescence photobleaching recovery (FPR) experiments are usually interpreted in terms of the translational motions of a fluorescently labeled species, rotational motions can also modulate recovery through the cosine-squared laws for dipolar absorption and emission processes. In a complex interacting system, translational and rotational contributions may both be simultaneously present. We show how these contributions can be separated in solution studies using an FPR setup in which (a) the linear polarization of the low-intensity observation beam and the high-intensity photobleaching pulse can be varied independently, and (b) all emitted fluorescent photons are counted equally. The fluorescence recovery signal obtained with the observation beam polarized at the magic angle, 54.7 degrees, from the bleach polarization direction is independent of label orientation, whereas the anisotropy function formed from a combination of parallel and perpendicular polarizations isolates the orientational recovery. The anisotropy function is identical to that in fluorescence correlation spectroscopy and, for rigid-body rotational diffusion, can be expressed as a sum of five exponential terms.  相似文献   

9.
The kinetics of association of Escherichia coli 30S and 50S ribosomal subunits have been carried out as a function of temperature after a magnesium jump from 1.5 to 3 mM. Turbidimetric recordings combined with a stopped-flow apparatus were used to follow the kinetics. The data show that the rates of formation and dissociation of the 70S particles at 3 mM Mg2+ and +25 degrees C were, respectively: k2 = 10(5) M-1 s-1, k1 = 4,5 X 10(-3) s-1; lowering the temperature decreases the rate constants with activation energies equal to E2 = 7.5 kcal/mol, E1 = 26.5 kcal/mol and enhances the association equilibrium towards the 70S species with an enthalpy change (delta H degrees assoc = -19.9 kcal/mol) dominant over the entropy change (delta S degrees assoc = -33 cal/(deg mol)). These thermodynamic parameters were compared to those obtained from studies on the interactions of codon-anticodon in yeast phenylalanine transfer RNA as well as of ribooligonucleotides. The kinetic and thermodynamic data are shown to be consistent with 16S-23S RNA interaction.  相似文献   

10.
C Tesi  F Travers  T Barman 《Biochemistry》1990,29(7):1846-1852
The initial steps of actomyosin subfragment 1 (acto-S1) ATPase (dissociation and binding of ATP) were studied at -15 degrees C with 40% ethylene glycol as antifreeze. The dissociation kinetics were followed by light scattering in a stopped-flow apparatus, and the binding of ATP was followed by the ATP chase method in a rapid-flow quench apparatus. The data from the chase experiments were fitted to E + ATP in equilibrium (K1) E.ATP----(k2) E*ATP, where E is acto-S1 or S1. The kinetics of the binding of ATP to acto-S1 were sensitive to the degree of saturation of the actin with S1. There was a sharp transition with actin nearly saturated with S1: when the S1 to actin ratio was low, the kinetics were fast (K1 greater than 300 microM, k2 greater than 40 s-1); when it was high, they were slow (K1 = 14 microM, k2 = 2 s-1). With S1 alone K1 = 12 microM and k2 = 0.07 S-1. With acto heavy meromyosin (acto-HMM) the binding kinetics were the same as with saturated acto-S1, regardless of the HMM to actin ratio. The dissociation kinetics were independent of the S1 to actin ratio. Saturation kinetics were obtained with Kd = 460 microM and kd = 75 S-1. The data for the saturated acto-S1 could be fitted to a reaction scheme, but for lack of structural information the abrupt dependence of the ATP binding kinetics upon the S1 to actin ratio is difficult to explain.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The kinetics of reduction of indigocarmine-dye-oxidized Fe protein of nitrogenase from Klebsiella pneumoniae (Kp2ox) by sodium dithionite in the presence and absence of MgADP were studied by stopped-flow spectrophotometry at 23 degrees C and at pH 7.4. Highly co-operative binding of 2MgADP (composite K greater than 4 X 10(10) M-2) to Kp2ox induced a rapid conformation change which caused the redox-active 4Fe-4S centre to be reduced by SO2-.(formed by the predissociation of dithionite ion) with k = 3 X 10(6) M-1.s-1. This rate constant is at least 30 times lower than that for the reduction of free Kp2ox (k greater than 10(8) M-1.s-1). Two mechanisms have been considered and limits obtained for the rate constants for MgADP binding/dissociation and a protein conformation change. Both mechanisms give rate constants (e.g. MgADP binding 3 X 10(5) less than k less than 3 X 10(6) M-1.s-1 and protein conformation change 6 X 10(2) less than k less than 6 X 10(3) s-1) that are similar to those reported for creatine kinase (EC 2.7.3.2). The kinetics also show that in the catalytic cycle of nitrogenase with sodium dithionite as reductant replacement of 2MgADP by 2MgATP occurs on reduced and not oxidized Kp2. Although the Kp2ox was reduced stoichiometrically by SO2-. and bound two equivalents of MgADP with complete conversion into the less-reactive conformation, it was only 45% active with respect to its ability to effect MgATP-dependent electron transfer to the MoFe protein.  相似文献   

12.
3,4-Dihydroxyphenylalanine (DOPA) is not a preferred substrate of Rhus vernicifera laccase, as rate constants for the anaerobic reduction of the type 1 cupric atom by L-DOPA (6.3 X 10(1) M-1 s-1), D-DOPA (2.6 X 10(1) M-1 s-1), and L-DOPA methyl ester (2.6 X 10(1) M-1 s-1) are considerably smaller than k1 (catechol) (7 X 10(2) M-1 s-1) and rate constants characteristic of numerous other nonphysiological organic substrates (25 degrees C, pH 7.0, I = 0.5 M). The reactions of DOPA derivatives with laccase are unique, however, in that a two-term rate law pertains: kobsd = k0 + k1[phenol]; k0(L-DOPA) = 7 X 10(-2) s-1. The reactivities of other catechol derivatives (pyrogallol, gallic acid, and methyl gallate) with laccase type 1 copper were also examined.  相似文献   

13.
We have used the fluorescence photobleaching recovery technique to study the dependence on oxygen tension of the lateral mobility of fluorescently labeled band 3, the phospholipid analogue fluorescein phosphatidylethanolamine, and glycophorins in normal red blood cell membranes. Band 3 protein and sialic acid moieties on glycophorins were labeled specifically with eosin maleimide and fluorescein thiosemicarbazide, respectively. The band 3 diffusion rate increased from 1.7 x 10(-11) cm2 s-1 to 6.0 x 10(-11) cm2 s-1 as oxygen tension was decreased from 156 to 2 torr, and a further increase to 17 x 10(-11) cm2 s-1 occurred as oxygen tension was decreased from 2 to 0 torr. The fractional mobility of band 3 decreased from 58 to 32% as oxygen tension was decreased from 156 to 0 torr. The phospholipid diffusion coefficient remained constant as oxygen tension was decreased from 156 to 20 torr, but increased from 2.3 x 10(-9) cm2 s-1 to 7.1 x 10(-9) cm2 s-1 as oxygen tension was decreased from 20 to 0 torr. Neither the diffusion coefficient nor the fractional mobility of glycophorins changed significantly at low oxygen tension. Under non-bleaching excitation conditions, intensities of fluorescence emission were identical for oxygenated and deoxygenated eosin-labeled RBCs. Deoxygenated eosin-labeled RBCs required 160-fold greater laser intensities than did oxygenated RBCs to achieve comparable extents of photobleaching, however. Oxygen seems to act as a facilitator of fluorophore photobleaching and may thereby protect the fluorescently labeled red cell membrane from photodamage.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
In the original theoretical development of fluorescence photobleaching recovery with circular or Gaussian laser intensity profiles (Axelrod et al., 1976, Biophys. J.) the bleaching process is assumed to obey first order kinetics in the fluorescent probe. While this is reasonable in most cases where oxygen participates in the photolysis reaction, some processes may obey second order kinetics in the fluorophore concentration due to dimerization. Accordingly, we present here an analysis of the fluorescence recovery when the photobleaching process is taken to be second order in the probe. Analytical solutions for small bleaching levels indicate that the fluorescence recovery curve is very similar to that measured following a bleaching process first order in the probe. Numerical solutions for moderate bleaching levels show that the recovery is qualitatively similar, but quantitatively different. Because the shape of the recovery curve provides no evidence as to the order of photobleaching, we recommend continued use of the previous theoretical analysis. However, it must be borne in mind that the diffusion coefficient is increasingly underestimated as the extent of photobleaching is increased. The true diffusion coefficient is obtained in the limit of small levels of photobleaching. Estimates of the fractional recovery are not affected by this approach.  相似文献   

15.
Fluorescence recovery after photobleaching (FRAP) is a widely used tool for estimating mobility parameters of fluorescently tagged molecules in cells. Despite the widespread use of confocal laser scanning microscopes (CLSMs) to perform photobleaching experiments, quantitative data analysis has been limited by lack of appropriate practical models. Here, we present a new approximate FRAP model for use on any standard CLSM. The main novelty of the method is that it takes into account diffusion of highly mobile molecules during the bleach phase. In fact, we show that by the time the first postbleach image is acquired in a CLSM a significant fluorescence recovery of fast-moving molecules has already taken place. The model was tested by generating simulated FRAP recovery curves for a wide range of diffusion coefficients and immobile fractions. The method was further validated by an experimental determination of the diffusion coefficient of fluorescent dextrans and green fluorescent protein. The new FRAP method was used to compare the mobility rates of fluorescent dextrans of 20, 40, 70, and 500 kDa in aqueous solution and in the nucleus of living HeLa cells. Diffusion coefficients were lower in the nucleoplasm, particularly for higher molecular weight dextrans. This is most likely caused by a sterical hindrance effect imposed by nuclear components. Decreasing the temperature from 37 to 22 degrees C reduces the dextran diffusion rates by approximately 30% in aqueous solution but has little effect on mobility in the nucleoplasm. This suggests that spatial constraints to diffusion of dextrans inside the nucleus are insensitive to temperature.  相似文献   

16.
Rate constants for most of the steps of the reaction cycle of the sarcoplasmic reticulum calcium-ATPase are similar or identical with Ca2+ or Sr2+ as the transported ions in spite of the large differences in the size and affinity of Ca2+ and Sr2+ (5 mM MgCl2, 100 mM KCl, pH 7.0, 25 degrees C). Phosphorylation of cE.Sr2 and cE.Ca2 by ATP occurs with kp = 220-235 s-1, whereas phosphorylation of E.ATP+Ca2+ or Sr2+ is consistent with kb = 50-70 s-1. Hydrolysis of E approximately P.Sr2 and E approximately P.Ca2 occurs with kt = 20 s-1, and the addition of 7 mM ADP to E approximately P.Sr2 or to E approximately P.Ca2 gives a burst of approximately 43% dephosphorylation, followed by dephosphorylation with k = 46 s-1. However, one Sr2+ ion dissociates from cE.Sr2 and from cE.ATP.Sr2 with k congruent to 120 s-1, whereas one Ca2+ ion dissociates from cE.Ca2 with k = 38 s-1 and from cE.ATP.Ca2 with k = 80 s-1.  相似文献   

17.
The influence of polyhydric alcohols and carbohydrates on the thermostability, i.e., the heat inactivation kinetics, of Bacillus licheniformis alpha-amylase was studied in the temperature range 96 degrees to 130 degrees C. High concentrations (from 9 to 60 weight percent) of glycerol, sorbitol, mannitol, sucrose, or starch can markedly decrease the inactivation rate constant, k, and in the studied cases, this stabilizing effect grows stronger with increasing additive concentration. Statements about stabilization should, however, be specified carefully with respect to temperature, because E(A) is mostly altered likewise. For dissolved enzyme E(A) was almost always decreased in the presence of polyol or carbohydrate, whereas for immobilized enzyme it was augmented in each studied instance. The inactivation of dissolved enzyme can, in all the studied cases, be adequately described as a first-order process. Immobilized enzyme, however, shows biphasic then first-order inactivation kinetics, depending on the additive concentration and temperature.  相似文献   

18.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
EB Brown  ES Wu  W Zipfel    WW Webb 《Biophysical journal》1999,77(5):2837-2849
Multiphoton fluorescence photobleaching recovery (MP-FPR) is a technique for measuring the three-dimensional (3D) mobility of fluorescent molecules with 3D spatial resolution of a few microns. A brief, intense flash of mode-locked laser light pulses excites fluorescent molecules via multiphoton excitation in an ellipsoidal focal volume and photobleaches a fraction. Because multiphoton excitation of fluorophores is intrinsically confined to the high-intensity focal volume of the illuminating beam, the bleached region is restricted to a known, three-dimensionally defined volume. Fluorescence in this focal volume is measured with multiphoton excitation, using the attenuated laser beam to measure fluorescence recovery as fresh unbleached dye diffuses in. The time course of the fluorescence recovery signal after photobleaching can be analyzed to determine the diffusion coefficient of the fluorescent species. The mathematical formulas used to fit MP-FPR recovery curves and the techniques needed to properly utilize them to acquire the diffusion coefficients of fluorescently labeled molecules within cells are presented here. MP-FPR is demonstrated on calcein in RBL-2H3 cells, using an anomalous subdiffusion model, as well as in aqueous solutions of wild-type green fluorescent protein, yielding a diffusion coefficient of 8.7 x 10(-7) cm(2)s(-1) in excellent agreement with the results of other techniques.  相似文献   

20.
Kinetics and mechanism of bilirubin binding to human serum albumin   总被引:3,自引:0,他引:3  
The kinetics of bilirubin binding to human serum albumin at pH 7.40, 4 degrees C, was studied by monitoring changes in bilirubin absorbance. The time course of the absorbance change at 380 nm was complex: at least three kinetic events were detected including the bimolecular association (k1 = 3.8 +/- 2.0 X 10(7) M-1 S-1) and two relaxation steps (52 = 40.2 +/- 9.4 s-1 and k3 = 3.8 +/- 0.5 s-1). The presence of the two slow relaxations was confirmed under pseudo-first order conditions with excess albumin. Curve-fitting procedures allowed the assignment of absorption coefficients to the intermediate species. When the bilirubin-albumin binding kinetics was observed at 420 nm, only the two relaxations were seen; apparently the second order association step was isosbestic at this wavelength. The rate of albumin-bound bilirubin dissociation was measured by mixing the pre-equilibrated human albumin-bilirubin complex with bovine albumin. The rate constant for bilirubin dissociation measured at 485 nm was k-3 = 0.01 s-1 at 4 degrees C. A minimum value of the equilibrium constant for bilirubin binding to human albumin determined from the ratio k1/k-3 is therefore approximately 4 X 10(9) M-1.  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