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.
To probe the functional role of tryptophan 49 in human antithrombin III, a mutant antithrombin, W49K, has been expressed in baby hamster kidney cells. The mutation reduces the affinity for heparin pentasaccharide by 1.8 kcal mol-1 but does not alter the heparin enhancement of the rate of factor Xa inhibition. 1H NMR spectra of W49K antithrombin show that the structure of the protein and the mode of heparin binding appear to be unaltered by the mutation, although tryptophan 49 is perturbed by heparin binding. 19F NMR spectra of 6-fluorotryptophan-substituted antithrombin show that tryptophan 49 is in a solvent-exposed environment. The heparin-induced fluorescence enhancement of W49K antithrombin is significantly different from that of wild-type antithrombin. Pentasaccharide induces only a 24% enhancement of antithrombin fluorescence, while high affinity heparin induces an enhancement of 40%. The results indicate that tryptophan 49 is probably a heparin contact residue but can be mutated without altering the remaining heparin-antithrombin interactions or the heparin-induced conformational change and resultant activation toward Factor Xa. Hydrophobic as well as charge interactions are thus probably involved in the specificity of the antithrombin-heparin pentasaccharide interaction. The lower fluorescence enhancements suggest that the heparin-induced 40% fluorescence enhancement used as the hallmark of activating heparin species is not the best indicator of the structural change in antithrombin that results in enhancement of the rate of proteinase inhibition.  相似文献   

2.
Studies of the iodide ion quenching of the intrinsic fluorescence of Concanavalin A indicate that 50% of the tryptophyl fluorescence originates from exposed residues. This agrees with the X-ray crystallographic determination that two of the four tryptophan residues in a Concanavalin A monomer are on the surface. Previous studies have indicated that conformational changes induced by sugar binding alter the environment of aromatic residues. The present investigation finds that neither the specific binding of alpha-methyl-D-mannoside nor alteration of the Concanavalin A quaternary structure changes the number or accessibility of the solvent-exposed tryptophan residues. It therefore appears that the major conformational transitions in Concanavalin A do not affect steric access to the surface tryptophans and the effects previously observed may be ascribed to structurally internal tryptophan residues.  相似文献   

3.
The presence of two unfolding domains in antithrombin III during its denaturation in guanidinium chloride has previously been reported (Villanueva, G. B., and Allen, N. (1983) J. Biol. Chem. 258, 11010-11013). In the present work, we report the results of refolding studies on antithrombin III. Circular dichroism and intrinsic fluorescence studies have demonstrated that the first unfolding domain of low stability (midpoint at 0.7 M guanidinium chloride) is irreversible upon renaturation, whereas the second unfolding domain (midpoint at 2.3 M guanidinium chloride) is reversible. The intermediate form of antithrombin III, termed AT-IIIR, which has lost the structural features of the first domain was investigated. Clotting assays and electrophoretic analyses showed that AT-IIIR had lost 60% of heparin cofactor activity but was still capable of forming sodium dodecyl sulfate-stable complexes with thrombin. Although certain regions of this molecule do not refold to the conformation of native antithrombin III, the tryptophan residues refold to a conformation identical with the native state. This was demonstrated by fluorescence quenching, solvent perturbation, and chemical modification studies. However, the tryptophan-ascribed fluorescence enhancement and absorption difference spectrum which occur when heparin binds to antithrombin III are reduced by 70%. On the basis of these data, the binding of heparin to antithrombin III is interpreted in terms of a two-step mechanism. The primary binding occurs in the region without tryptophan and is followed by a secondary conformational rearrangement which affects the tryptophan environment. The mechanism of the binding of heparin and antithrombin III has been previously studied by kinetic methods, and the data also support a two-step mechanism. The agreement of these two studies employing entirely different approaches to the same problem lends support to the validity of this postulated mechanism.  相似文献   

4.
The acidic transition of porcine plasminogen, prepared by affinity chromatography, was studied by non-destructive methods. These methods are based on the analysis of the behaviour of the tryptophyls under various conditions. The perturbation of the absorption and emission spectra by pH or temperature and the dynamic quenching of the intrinsic fluorescence are used to obtain information on structural changes which affect the environment of these residues. It is shown that by decreasing pH the fluorescence emission spectra are shifted toward the long wavelengths, with a broadening of the fluorescence band. The same effect can be obtained at constant pH by heating the protein solution. In order to analyze these phenomena, it is assumed that the fluorescence intensities at 355 nm and 328 nm reflect the proportion of the tryptophans which are exposed to the solvent, and buried, respectively. The plot of the ratio of the fluorescence intensities at these wavelengths versus pH or temperature leads to a titration curve showing an unmasking of tryptophans. The proportion of exposed tryptophans is measured by the dynamic fluorescence quenching technique and the data analyzed according to Lehrer. The plot of the fraction of exposed tryptophyls versus pH also shows the unmasking of these chromophores. Thermal perturbation of a solution of plaminogen at neutral pH induces a difference absorption spectrum whose amplitudes at the maxima are proportional to the number of exposed aromatic residues. The comparison with a solution of fully denatured plasminogen in 6 M guanidium chloride, where all the tryptophyls are exposed, shows that the percentage of exposure is equal to 59%. This number is significantly higher than the percentage found by the fluorescence quenching technique (20%), indicating that some tryptophyls are located in crevices, exposed to the solvent but not to the iodide. At acidic pH the absorption difference spectra induced by thermal perturbation are not classical, since they show an inversion and a new band between 300 nm and 305 nm. This band is mentioned in the literature as a minor band of tryptophan which appears when this chromophore is located in an asymmetric environment. On plotting the maximum amplitude of these spectra obtained at acidic pH versus temperature, we obtain a curve indicating that two types of antagonistic interactions are involved in the perturbation of the chromophores spectra. The spectrophotometric titration of plasminogen gives classical absorption difference spectra. By plotting the maximum amplitude at 292 nm versus pH, we obtain a titration curve with an apparent pK of 2.9 units. This pK is acidic which respect to the pK value of a normal carboxyl. This low value can be due to a positively charged group in the neighbourhood of a carboxyl, which interacts with one or more chromophores. When the carboxyl becomes protonated, this positively charged group is free and available to perturb the environment of some chromophores...  相似文献   

5.
Corrected fluorescence excitation and emission spectra of human antithrombin III have been determined. The fluorescence observed originates almost entirely from tryptophan residues. Reduction of the disulfide bonds followed by carboxymethylation did not change the fluorometric properties of the protein. The binding of heparin to antithrombin III caused a marked fluorescence enhancement by about 30% of the intrinsic protein emission intensity. Various samples of heparin yielded different binding curves. Heparin fractionated by gel filtration seemed to be bound to two sites on antithrombin III with association constants of 0.6-10(6)m-1 and 0.2-10(6)M-1 respectively. Heparin, prepared by affinity chromatography on matrix-bound antithrombin III appeared to be bound to only one site with an association constant of 2.3-10(6)M-1. Under similar conditions heparin caused no increase of the intrinsic protein emission intensity when added to reduced and carboxymethylated antithrombin III. The implications of these findings are discussed.  相似文献   

6.
The conformation of native pertussis toxin has been investigated by secondary structure prediction and by circular dichroism, fluorescence and second-derivative ultraviolet absorption spectroscopy. The far-ultraviolet circular dichroic spectrum is characteristic of a protein of high beta-sheet and low alpha-helix content. This is also shown by an analysis of the circular dichroic spectrum with the Contin programme which indicates that the toxin possesses 53% beta-sheet, 10% alpha-helix and 37% beta-turn/loop secondary structure. Second-derivative ultraviolet absorption spectroscopy suggests that 34 tyrosine residues are solvent-exposed and quenching of tryptophan fluorescence emission has shown that 4 tryptophan residues are accessible to iodide ions. One of these tryptophans appears to be in close proximity to a positively charged side-chain, since only 3 tryptophans are accessible to caesium ion fluorescence quenching. When excited at 280 nm, the emission spectrum contains a significant contribution from tyrosine fluorescence, which may be a consequence of the high proportion (55%) of surface-exposed tyrosines. No changes in the circular dichroic spectra of the toxin were found in the presence of the substrate NAD. However, NAD did quench both tyrosine and tryptophan fluorescence emission but did not change the shape of the emission spectrum, or the accessibility of the tryptophans to either the ionic fluorescence quenchers or the neutral quencher acrylamide.  相似文献   

7.
Dusa A  Kaylor J  Edridge S  Bodner N  Hong DP  Fink AL 《Biochemistry》2006,45(8):2752-2760
The aggregation of the presynaptic protein alpha-synuclein is associated with Parkinson's disease (PD). The details of the mechanism of aggregation, as well as the cytotoxic species, are currently not well understood. alpha-Synuclein has four tyrosine and no tryptophan residues. We introduced a tyrosine to tryptophan mutation at position 39 to create an intrinsic fluorescence probe and allow additional characterization of the aggregation process. Y39W alpha-synuclein had similar fibrillation kinetics (2-fold slower), pH-induced conformational changes, and fibril morphology to wild-type alpha-synuclein. In addition to intrinsic Trp fluorescence, acrylamide quenching, fluorescence anisotropy, ANS binding, dynamic light scattering, and FTIR were employed to monitor the kinetics of aggregation. These biophysical probes revealed the significant population of two classes of oligomeric intermediates, one formed during the lag period of fibrillation and the other present at the completion of fibrillation. As expected for a natively unfolded protein, Trp 39 was highly solvent-exposed in the monomer and is solvent-exposed in the two oligomeric intermediates; however, it is partially, but not fully, buried in the fibrils. These observations demonstrate the utility of Trp fluorescence labeled alpha-synuclein and demonstrate the existence of an oligomeric intermediate that exists as a transient reservoir of alpha-synuclein for fibrillation.  相似文献   

8.
Chemical modifications have demonstrated that the ultraviolet difference spectrum produced when heparin interacts with antithrombin III is due primarily to changes in the tryptophan environment. This is based on the observation that this spectrum could be abolished by treatment of antithrombin III with dimethyl (2-hydroxy-5-nitrobenzyl) sulfonium bromide but not with tetranitromethane. The tryptophan-modified antithrombin III is still capable of binding to thrombin even when it has lost 85% of heparin cofactor activity. A marked decrease in reactivity of tryptophan residues is observed when modification is carried out in the presence of heparin. Evidence is presented that tryptophan is in the heparin binding site.  相似文献   

9.
A combination of intrinsic fluorescence and circular dichroic (CD) spectroscopy has been used to characterize the complexes formed between bovine retinal arrestin and heparin or phytic acid, two ligands that are known to mimic the structural changes in arrestin attending receptor binding. No changes in the CD spectra were observed upon ligand binding, nor did the degree of tryptophan fluorescence quenching change significantly in the complexes. These data argue against any large-scale changes in protein secondary or tertiary structure accompanying ligand binding. The change in tyrosine fluorescence intensity was used to determine the dissociation constants for the heparin and phytic acid complexes of arrestin. The only change observed was a saturable diminution of tyrosine fluorescence signal from the protein. For both ligands, the data suggest two distinct binding interactions with the protein—a high-affinity interaction with K d between 200 and 300 nM, and a lower affinity interaction with K d between 2 and 8 M. Study of collisional quenching of tyrosine fluorescence in free arrestin and the ligand-replete complexes indicates that 10 of the 14 tyrosine residues of the protein are solvent-exposed in the free protein; this value drops to between 5 and 6 solvent-exposed residues in the high-affinity complexes of the two ligands. These data suggest that ligand binding leads to direct occlusion of between 4 and 5 tyrosine residues on the solvent-exposed surface of the protein, but not to any large-scale changes in protein structure. The large activation energy previously reported to be associated with arrestin–receptor interactions may therefore reflect localized movements of the N- and C-termini of arrestin, which are proposed to interact in the free protein through electrostatic interactions. Binding of the anionic ligands heparin, phytic acid, or phosphorylated rhodopsin may compete with the C-terminus of arrestin for these electrostatic interactions, thus allowing the C-terminus to swing out of the binding region.  相似文献   

10.
The quenching of tryptophan fluorescence by N-bromosuccinamide, studied by the fluorescence stopped-flow technique, was used to compare the reactivities of tryptophan residues in protein molecules. The reaction of N-bromosuccinamide with the indole group of N-acetyltryptophanamide, a model compound for bound tryptophan, followed second-order kinetics with a rate constant of (7.8 +/- 0.8) . 10(5) dm3 . mol-1 . s-1 at 23 degrees C. The rate does not depend on ionic strength or on the pH near neutrality. The non-fluorescent intermediate formed from N-acetyltryptophanamide on the reaction with N-bromosuccinamide appears to be a bromohydrin compound. The second-order rate constant for fluorescence quenching of tryptophan in Gly-Trp-Gly by N-bromosuccinamide was very similar, (8.8 +/- 0.8) . 10(5) dm3 . mol-1 . s-1. Apocytochrome c has the conformation of a random coil with the single tryptophan largely exposed to the solvent. The rate constant for the fluorescence quenching of the tryptophan in apocytochrome c by N-bromosuccinamide was (3.7 +/- 0.3) . 10(5) dm3 . mol-1 . s-1. The fluorescence quenching by N-bromosuccinamide of the tryptophan residues incorporated in alpha-chymotrypsin at pH 7.0 showed three exponential terms from which the following rate constants were derived: 1.74 . 10(5), 0.56 . 10(5) and 0.11 . 10(5) dm3 . mol-1 . s-1. This protein is known to have eight tryptophan residues in the native state, six residues at the surface, and two buried. Three of the surface tryptophans have the indole rings protruding out of the molecule and may account for the fastest kinetic phase of the quenching process. The intermediate phase may be due to three surface tryptophans whose indole rings point inwards, and the slowest to the two interior tryptophan residues.  相似文献   

11.
The 40 S heterogeneous nuclear ribonucleoprotein (hnRNP) particles from HeLa cells reveal tryptophan fluorescence with a bi-exponential decay, indicating that only a few of the 'core' proteins contain tryptophan residues. The presence of tryptophan residues distinguishes hnRNP particles from nucleosomes, with which they otherwise share a number of properties. This difference, however, is not essential for protein-RNA binding, as the fluorescence decay remains unchanged when hnRNP particles are dissociated into protein and RNA. However, the Stern-Volmer quenching constant is doubled upon salt dissociation, i.e. tryptophan residues become more accessible to solvent. Thus tryptophan quenching is a useful parameter for monitoring protein-protein interactions in hnRNP particles.  相似文献   

12.
The intrinsic fluorescence of lauryl maltoside solubilized bovine heart cytochrome c oxidase has been determined to arise from tryptophan residues of the oxidase complex. The magnitude of the fluorescence is approximately 34% of that from n-acetyltryptophanamide (NATA). This level of fluorescence is consistent with an average heme to tryptophan distance of 30 A. The majority of the fluorescent tryptophan residues are in a hydrophobic environment as indicated by the fluorescence emission maximum at 328 nm and the differing effectiveness of the quenching agents: Cs+, I-, and acrylamide. Cesium was ineffective up to a concentration of 0.7 M, whereas quenching by the other surface quenching agent, iodide, was complex. Below 0.2 M, KI was ineffective whereas between 0.2 and 0.7 M 15% of the tryptophan fluorescence was found to be accessible to iodide. This pattern indicates that protein structural changes were induced by iodide and may be related to the chaotropic character of KI. Acrylamide was moderately effective as a quenching agent of the oxidase fluorescence with a Stern-Volmer constant of 2 M-1 compared with acrylamide quenching of NATA and the water-soluble enzyme aldolase having Stern-Volmer constants of 12 M-1 and 0.3 M-1, respectively. There was no effect of cytochrome c on the tryptophan emission intensity from cytochrome c oxidase under conditions where the two proteins form a tight, 1:1 complex, implying that the tryptophan residues near the cytochrome c binding site are already quenched by energy transfer to the homes of the oxidase. The lauryl maltoside concentration used to solubilize the enzyme did not affect the fluorescence of NATA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
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   

14.
Trifluoroacetamide was found to be a good quencher of tryptophan fluorescence, and the quenching was shown to proceed via both a dynamic and a static process. The respective quenching constants were determined by the measurement of the decrease of the fluorescence lifetime in the presence of the quencher. The static and the bimolecular rate quenching constants of N-acetyltryptophanamide are equal to 0.34 1·mol?1 and 1.9·109 1·mol?1·s?1, respectively. These values indicate that trifluoroacetamide is an efficient quencher of tryptophan fluorescence. This conclusion is also supported by a complete quenching of bovine serum albumin and wheat germ agglutinin fluorescence. In the case of lysozyme, trifluoroacetamide quenches the fluorescence of tryptophan residues which fluoresce with a maximum at 348 nm but not the buried tryptophan residues which fluoresce with a maximum at 333 nm. Trifluoroacetamide quenching of wheat germ agglutinin emission confirms the homogeneity and the high accessibility of emitting tryptophan residues, in agreement with a previous report (Privat, J.P. and Monsigny, M. (1975) Eur. J. Biochem. 60, 555–567). The tryptophan fluorescence decay of wheat germ agglutinin is biexponential even in the presence of the quencher; the static and bimolecular rate quenching constants are equal to 0.22 1·mol?1 and 092·109 1·mol?1·?1, respectively. In the presence of a specific lectin ligand, the methyldi-N,N′-trifluoroacetyl-β- chitobioside, the quenching of wheat germ agglutinin fluorescence involves a direct contact between tryptophan residues and trifluoroacetamido groups of the ligand and in contrast with the quenching induced by free trifluoroacetamide shows that the tryptophan fluorescence is not fully quenched.  相似文献   

15.
Cucurbitacin delta 23-reductase from Cucurbita maxima var. Green Hubbard fruit displays an apparent Mr of 32,000, a Stokes radius of 263 nm and a diffusion coefficient of 8.93 X 10(-7) cm2 X s-1. The enzyme appears to possess a homogeneous dimeric quaternary structure with a subunit Mr of 15,000. Two tryptophan and fourteen tyrosine residues per dimer were found. Emission spectral properties of the enzyme and fluorescence quenching by iodide indicate the tryptophan residues to be buried within the protein molecule. In the pH range 5-7, where no conformational changes were detected, protonation of a sterically related ionizable group with a pK of approx. 6.0 markedly influenced the fluorescence of the tryptophan residues. Protein fluorescence quenching was employed to determine the dissociation constants for binding of NADPH (Kd 17 microM), NADP+ (Kd 30 microM) and elaterinide (Kd 227 microM). Fluorescence energy transfer between the tryptophan residues and enzyme-bound NADPH was observed.  相似文献   

16.
The active site of an apoptotic enzyme caspase-3 was characterized by measuring the intrinsic fluorescence of two tryptophan residues. Temperature dependence of the intrinsic fluorescence, the energy homotransfer between the tryptophan residues, and the fluorescence quenching by tetrapeptide inhibitors were investigated by the fluorescence lifetime measurements. It has been observed that the fluorescence lifetimes of caspase-3 in complex with inhibitors were significantly shortened by the electron transfer process.  相似文献   

17.
The effect of ionizing radiation on tryptophan fluorescence of thymocyte and erythrocyte plasma membrane preparations was studied. The intensity of tryptophan fluorescence decreased after applying radiation doses up to 15 Gy. The radiosensitivity of thymocyte membranes appeared to be higher than that of the erythrocyte ghosts. Tryptophan radiolysis did not significantly contribute to the effects of radiation. The fraction of tryptophan residues accessible for quenching by I- decreased from 0.87 in the untreated membranes to 0.63 and 0.49 in membranes after doses of 10 and 250 Gy, respectively. The effective quenching constant and the tryptophan fluorescence polarization increased after irradiation. The mechanisms producing these radiation-induced changes are discussed.  相似文献   

18.
The acrylamide-quenching patterns of the intrinsic tryptophan fluorescence of six cold-soluble monoclonal immunoglobulin M (IgM) and two monoclonal IgM proteins possessing cryoglobulin properties (abnormal cold insolubility) have been compared. Static and dynamic components of quenching have been resolved by a modified form of the Stern-Volmer relationship. The unusual observation of static quenching seen with the multitryptophan containing IgM is determined to be a consequence of essentially homogeneous indole fluorescence arising from conserved tryptophan residues within each homologous immunoglobulin domain. Although the static component of the quenching of the two IgM cryoimmunoglobulins examined is similar to that of the non-cryoimmunoglobulin, IgM, some of the cryoglobulin's tryptophan residues appear to be more kinetically exposed to acrylamide than the tryptophans in the non-cryoglobulin IgM. An unusually large negative entropy of activation observed for the quenching process of both cryoimmunoglobulins suggests some abnormality in the dynamic (flexibility) properties of these proteins.  相似文献   

19.
Quenching of tryptophan fluorescence of maize and wheat NADP-malic enzyme by KI and acrylamide was studied after denaturating proteins with guanidine hydrochloride, and subjecting them to different pH values or temperatures. Protein unfolding by guanidine hydrochloride resulted in a red shift of the fluorescence spectrum, providing further support for the motion that several of the tryptophan residues evolved from an apolar to a polar environment. Protein denaturation was accompanied by an increase in the effective dynamic quenching constant values and by loss of the enzyme's activities. Thermal denaturation gave results consistent with the ones observed for chemical denaturation suggesting that a putative intermediate is involved in the denaturation process. Finally, exposure of both enzymes at various pH values allowed us to infer the number of accessible tryptophan residues in the different oligomeric conformations. The results suggest that the aggregation process seems to be different for each enzyme. Thus, as the maize enzyme associated from monomer to tetramer, one tryptophan residue would change from a polar to an apolar environment, while the association of the wheat enzyme would cause that two tryptophan residues to be excluded from quenching. Hitherto, quenching of the tryptophan fluorescence provides a good tool for studying conformational changes of proteins. The future availability of the crystal structures of plant NADP-malic enzymes will offer a good validation point for our model and the technology used.  相似文献   

20.
The tryptophan residues of two forms of pig heart lipoamide dehydrogenase (LD(I) and LD(II] were investigated fluorometrically. The tryptophan contents of LD(I) and LD(II) determined by the fluorescence method were 3 mol and 2 mol per mol of FAD, respectively. These values were in good agreement with those found by the MCD method. The microenvironments of the tryptophan residues were investigated by fluorescence quenching titration with acrylamide. The tryptophan residues of both enzymes were in heterogeneous microenvironments, and CD spectra showed some differences between these microenvironments in the two enzymes. Energy transfer from tryptophan residues to bound FAD was equally efficient in the two enzymes. It seems probable that the three tryptophan residues in LD(I) are all in different microenvironments, but that two of them are in microenvironments almost identical to those of the corresponding residues in LD(II).  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