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 46 毫秒
1.
Upon gradually heating a particular mutant of the flavoprotein NADH peroxidase, it was found from the peculiar time-resolved fluorescence anisotropy pattern of the flavin prosthetic group (FAD) that, at elevated temperature, FAD is released from the tetrameric enzyme. Since in this case a mixture of free and enzyme-bound FAD contributes to the time-dependent fluorescence anisotropy, its analysis can only be accomplished by an associative fitting model, in which specific fluorescence lifetimes of both species are linked to specific correlation times. In this letter the general approach to the associative polarized fluorescence decay analysis is described. The procedure can be used for other flavoproteins to determine the temperature at which the onset of thermal denaturation will start, leading to release of the flavin prosthetic group. Received: 20 November 1998 / Revised version: 6 April 1999 / Accepted: 8 April 1999  相似文献   

2.
Time-resolved polarized fluorescence spectroscopy has been applied to the bound FAD in the structurally related flavoproteins lipoamide dehydrogenase from Azotobacter vinelandii (LipDH-AV) and glutathione reductase (GR) from human erythrocytes. The fluorescence parameters as obtained from the maximum entropy analysis differ considerably in both enzymes, reflecting the unique properties of the flavin microenvironment. Three conformational substates are revealed in LipDH-AV and five in GR. Almost 90% of the population of GR molecules has a fluorescence lifetime in the order of 30 ps which originates from efficient exciplex formation with Tyr197. Equilibrium fluctuations between conformational substates are observed for LipDH-AV on a nanosecond time scale in the temperature range 277-313 K. Interconversion between conformational substates in GR is slow, indicating that large activation barriers exist between the states. In agreement with these results, a model is postulated which ascribes a role in catalysis to equilibrium fluctuations between conformational substates in GR and LipDH-AV. From time-resolved fluorescence anisotropy as a function of temperature, distinction can be made between flavin reorientational motion and interflavin energy transfer. In both proteins intersubunit energy transfer between the prosthetic groups is observed. Furthermore, it is revealed that only the flavin in glutathione reductase exhibits rapid restricted reorientational motion. Geometric information concerning the relative orientation and distance of the flavins can be extracted from the parameters describing the energy-transfer process. The obtained spatial arrangement of the flavins is in excellent agreement with crystallographic data.  相似文献   

3.
Flavin reductases use flavins as substrates and are distinct from flavoenzymes which have tightly bound flavins. The reduced flavin can serve to reduce ferric complexes and iron proteins. In Escherichia coli, reactivation of ribonucleotide reductase is achieved by reduced flavins produced by flavin reductase. The crystal structure of E. coli flavin reductase reveals that the enzyme structure is similar to the structures of the ferredoxin reductase family of flavoproteins despite very low sequence similarities. The main difference between flavin reductase and structurally related flavoproteins is that there is no binding site for the AMP moiety of FAD. The direction of the helix in the flavin binding domain, corresponding to the phosphate binding helix in the flavoproteins, is also slightly different and less suitable for phosphate binding. Interactions for flavin substrates are instead provided by a hydrophobic isoalloxazine binding site that also contains a serine and a threonine, which form hydrogen bonds to the isoalloxazine of bound riboflavin in a substrate complex.  相似文献   

4.
Malcolm Dixon 《BBA》1971,226(2):269-284
1. The specificity of flavoproteins towards acceptors has been rather neglected, but an attempt is here made to construct a comparative table of acceptor specificities of those flavoprotein enzymes for which data exist.

2. The acceptor specificity of reduced flavin groups, when combined with apoenzyme proteins, is quite different from that of the same flavin groups in the free state (see Part II). Free flavins react very rapidly with a wide range of acceptors, but the same groups combined as flavoproteins have a severely restricted range of action.

3. There are remarkable differences between different flavoproteins. Nearly every flavoprotein fails altogether to react with at least one, and often several, of the acceptors, giving a specificity pattern which is different in each case. There seems to be no general acceptor for flavoproteins.

4. The effect of combination of a flavin with a particular apoenzyme is to inhibit specifically the reaction of the flavin with particular acceptors with which it would react very rapidly in the absence of the apoenzyme.

5. Each apoenzyme produces its own distinctive pattern of inhibitions. The degree of inhibition is often very high; the table shows over 50 cases of specific inhibitions that are essentially complete. Some of these are very difficult to explain.

6. There is no obvious parallelism between any acceptor and any other in its pattern of reactivity with a series of different flavoproteins.

7. In a few cases combination with apoenzyme specifically accelerates the reaction of the flavin with particular acceptors, so that the flavoprotein is oxidized faster than the free flavin.

8. Possible correlations are discussed between the effects of apoenzymes on the reactivity of flavins with acceptors and a number of special known features of different apoenzymes, but no adequate explanation of the differences in specificity has emerged.

9. In view of the interesting nature of the effects, a plea is made for a more intensive study of the acceptor side of flavoprotein specificity.  相似文献   


5.
W S Kunz 《FEBS letters》1986,195(1-2):92-96
The different flavoproteins contributing to flavin fluorescence of isolated rat liver mitochondria have distinct excitation and emission spectra. The NAD-linked flavin component was identified as alpha-lipoamide dehydrogenase, while the non-NAD-linked component was found to be electron transfer flavoprotein. The differences in excitation and emission properties of the mitochondrial flavoproteins permit selective recording of their redox state changes in isolated mitochondria.  相似文献   

6.
Comparative investigation of different mitochondrial oxidative metabolism inhibitors action on NAD(P)H and flavoproteins fluorescence intensity of minimal transformed 3T3 NIH mouse fibroblasts and rat HTC hepatoma cells was made. Principle differences were shown between these cells in oxidized flavoproteins fluorescence intensity changes under the action of used inhibitors. It is suggested that the unusual HTC hepatoma cells flavin fluorescence intensity increase is connected with the oxidation of unidentified flavin-containing component functionally attached to mitochondrial respiratory chain.  相似文献   

7.
Light-mediated reduction of flavoproteins with flavins as catalysts   总被引:3,自引:0,他引:3  
It has been found that small amounts of free flavins greatly accelerate the photochemical reduction of flavoproteins both to the radical and fully reduced oxidation states. This catalytic effect has been shown to be due to the rapid photochemical reduction of the free flavin to its fully reduced state, followed by its reaction with the flavoprotein to yield flavoprotein radical and by its reaction with flavoprotein radical to yield fully reduced flavoprotein. Evidence is presented that the same route may occur with flavoproteins in the absence of added flavins. In this case the photoreduction is mediated by the small equilibrium concentration of free flavin coenzyme present in a flavorprotein solution. Hence, it is suggested that flavoprotein reduction with EDTA as photosubstrate does not involve an excited state of the holoprotein, nor contact of EDTA with the enzyme, but exchange of electrons between enzyme flavin and free reduced flavin.  相似文献   

8.
Analysis of fluorescence decay data for probes incorporated into model or biological membranes invariably requires fitting to more than one decay time even though the same probe exhibits nearly single-exponential decay in solution. The parinaric acids (cis and trans) are examples of this. Data are presented for both parinaric acid isomers in dimyristoylphosphatidylcholine membranes collected to higher precision than normally encountered, and the fluorescence decays are shown to be best described by a smooth distribution of decay times rather than by a few discrete lifetimes. The temperature dependence of the fluorescence decay reveals a clear shift in the distribution to longer lifetimes associated with the membrane phase transition at 23.5 degrees C. The physical significance is that fluorescence lifetime measurements appear to reflect a physical process with a distribution of lifetimes rather than several distinct physical processes.  相似文献   

9.
The fluorescence signal of flavoproteins of rat liver mitochondria was investigated to determine the respective contributions of the various flavoenzymes. About 50% of the overall signal were found to be NAD-linked and caused by alpha-lipoamide dehydrogenase flavin (Em7.4 = -283 mV). Roughly 25% were due to a flavoprotein reducible in a non-NAD-linked reaction. This fluorescent flavoenzyme (Em7.4 = -52 mV) has been tentatively identified as a flavoprotein of the fatty-acid-oxidizing system, most probably the electron transfer flavoprotein. The remaining 25% of the signal are accounted for by flavoenzymes which are reducible by dithionite only. These flavoenzymes were not involved in the flavoprotein fluorescence alterations accompanying changes in electron flow through the respiratory chain. Contributions of other mitochondrial flavoproteins such as succinate dehydrogenase, NADH dehydrogenase, alpha-glycerophosphate dehydrogenase, proline dehydrogenase, and choline oxidase, to the overall flavin fluorescence signal of isolated rat liver mitochondria can be neglected.  相似文献   

10.
Time-resolved polarized flavin fluorescence was used to study the active site dynamics of Escherichia coli glutathione reductase (GR). Special consideration was given to the role of Tyr177, which blocks the access to the NADPH binding-site in the crystal structure of the enzyme. By comparing wild-type GR with the mutant enzymes Y177F and Y177G, a fluorescence lifetime of 7 ps that accounts for approximately 90% of the fluorescence decay could be attributed to quenching by Y177. Based on the temperature invariance for this lifetime, and the very high quenching rate, electron transfer from Y177 to the light-excited isoalloxazine part of flavin adenine dinucleotide (FAD) is proposed as the mechanism of flavin fluorescence quenching. Contrary to the mutant enzymes, wild-type GR shows a rapid fluorescence depolarization. This depolarization process is likely to originate from a transient charge transfer interaction between Y177 and the light-excited FAD, and not from internal mobility of the flavin, as has previously been proposed. Based on the fluorescence lifetime distributions, the mutants Y177F and Y177G have a more flexible protein structure than wild-type GR: in the range of 223 K to 277 K in 80% glycerol, both tyrosine mutants mimic the closely related enzyme dihydrolipoyl dehydrogenase. The fluorescence intensity decays of the GR enzymes can only be explained by the existence of multiple quenching sites in the protein. Although structural fluctuations are likely to contribute to the nonexponential decay and the probability of quenching by a specific site, the concept of conformational substates need not be invoked to explain the heterogeneous fluorescence dynamics.  相似文献   

11.
The resonance Raman (RR) spectra of FMN, FAD, FAD in D2O, and 7,8-dimethyl-1, 10-ethyleneisoalloxazinium perchlorate have been obtained by employing KI as a collisional fluorescence-quenching agent. The spectra are very similar to those obtained recently by using the CARS technique to eliminate fluorescence. Spectra have also been obtained for several species in which flavin is known to fluoresce only weakly. We report RR spectra of protonated FMN, FMN semiquinone cation, the general fatty acyl-CoA dehydrogenase, and two "charge-transfer" complexes of fatty acyl-CoA dehydrogenase. Tentative assignment of several vibrational bands can be made on the basis of our flavin spectra. RR spectra of fatty acyl-CoA and its complexes are consistent with the previous hypothesis that visible spectral shifts observed during formation of acetoacetyl-CoA and crotonyl-CoA complexes of fatty acyl-CoA dehydrogenase result from charge-transfer interactions in which the ground state is essentially nonbonding as opposed to interactions in which complete electron transfer occurs to form FAD semiquinone. The only significant change in the RR spectrum of FAD on binding to enzyme occurs in the 1250-cm-1 region of the spectrum, a region associated with delta N--H of N-3. The position of this band in fatty acyl-CoA dehydrogenase and the other flavoproteins studied to date is discussed in terms of hydrogen bonding between flavin and protein.  相似文献   

12.
The authors demonstrate that the fluorescence lifetime of certain fluorescent labels is a useful parameter to detect affinity binding between biotin and streptavidin, as well as between biotinylated bovine serum albumin and streptavidin. The assay is performed in a microplate format, and lifetimes are determined using dye laser-induced fluorescence. Four fluorescent labels are presented that undergo a significant change in their lifetime upon affinity binding. The scheme, referred to as the fluorescence lifetime affinity assay, has several attractive features in that it requires single labeling only, represents a homogeneous assay, allows each of the 2 binding partners to be labeled, and is compatible with the standard microwell formats used in high-throughput screening.  相似文献   

13.
The structural dynamics of bovine erythrocyte Cu, Zn superoxide dismutase (BSOD) was studied by time-resolved fluorescence spectroscopy. BSOD is a homodimer containing a single tyrosine residue (and no tryptophan) per subunit. Frequency-domain fluorometry revealed a heterogeneous fluorescence decay that could be described with a Lorentzian distribution of lifetimes. The lifetime distribution parameters (center and width) were markedly dependent on temperature. The distribution center (average lifetime) displayed Arrhenius behavior with an Ea of 4.2 kcal/mol, in contrast with an Ea of 7.4 kcal/mol for the single-exponential decay of L-tyrosine. This indicated that thermal quenching of tyrosine emission was not solely responsible for the effect of temperature on the lifetimes of BSOD. The distribution width was broad (1 ns at 8 degrees C) and decreased significantly at higher temperatures. Furthermore, the width of the lifetime distribution increased in parallel to increasing viscosity of the medium. The combined effects of temperature and viscosity on the fluorescence decay suggest the existence of multiple conformational substrates in BSOD that interconvert during the excited-state lifetime. Denaturation of BSOD by guanidine hydrochloride produced an increase in the lifetime distribution width, indicating a larger number of conformations probed by the tyrosine residue in the denatured state. The rotational mobility of the tyrosine in BSOD was also investigated. Analysis of fluorescence anisotropy decay data enabled resolution of two rotational correlation times. One correlation time corresponded to a fast (picosecond) rotation that contributed 62% of the anisotropy decay and likely reported local mobility of the tyrosine ring. The longer correlation time was 50% of the expected value for rotation of the whole (dimeric) BSOD molecule and appeared to reflect segmental motions in the protein in addition to overall tumbling. Comparison between rotational correlation times and fluorescence lifetimes of BSOD indicates that the heterogeneity in lifetimes does not arise from mobility of the tyrosine per se, but rather from dynamics of the protein matrix surrounding this residue which affect its fluorescence decay.  相似文献   

14.
Time-resolved flavin fluorescence anisotropy studies on glutathione reductase (GR) have revealed a remarkable new phenomenon: wild-type GR displays a rapid process of fluorescence depolarization, that is absent in mutant enzymes lacking a nearby tyrosine residue that blocks the NADPH-binding cleft. Fluorescence lifetime data, however, have shown a more rigid active-site structure for wild-type GR than for the tyrosine mutants. These results suggest that the rapid depolarization in wild-type GR originates from an interaction with the flavin-shielding tyrosine, and not from restricted reorientational motion of the flavin. A novel mechanism of fluorescence depolarization is proposed that involves a transient charge-transfer complex between the tyrosine and the light-excited flavin, with a concomitant change in the direction of the emission dipole moment of the flavin. This interaction is likely to result from side-chain relaxation of the tyrosine in the minor fraction of enzyme molecules in which this residue is in an unsuitable position for immediate fluorescence quenching at the moment of excitation. Support for this mechanism is provided by binding studies with NADP+ and 2'P-5'ADP-ribose that can intercalate between the flavin and tyrosine and/or block the latter. Fluorescence depolarization analyses as a function of temperature and viscosity confirm the dynamic nature of the process. A comparison with fluorescence depolarization effects in a related flavoenzyme indicates that this mechanism of flavin fluorescence depolarization is more generally applicable.  相似文献   

15.
The conformational dynamics of wild-type Escherichia coli thioredoxin reductase (TrxR) and the mutant enzyme C138S were studied by ultrafast time-resolved fluorescence of the flavin cofactor in combination with circular dichroism (both in the flavin fingerprint and far-UV regions) and steady-state fluorescence and absorption spectroscopy. The spectroscopic data show two conformational states of the enzyme (named FO and FR), of which the physical characteristics differ considerably. Ultrafast fluorescence lifetime measurements make it possible to distinguish between the two different populations: Dominant picosecond lifetimes of approximately 1 ps (contribution 75%) and 7 ps (8%) are associated with the FO species in TrxR C138S. Long-lived fluorescence with two time constants in the range of 0.2-1 ns (total contribution 17%) originates from enzyme molecules in the FR conformation. The near absence of fast lifetime components in oxidized wild-type TrxR supports the idea of this enzyme being predominantly in the FR conformation. The emission spectrum of the FO conformation is blue-shifted with respect to that of the FR conformation. Because of the large difference in fluorescence characteristics, fluorescence measurements on time scales longer than 100 ps are fully determined by the fraction of enzyme molecules in the FR conformation. Binding of the thiol reagent phenyl mercuric acetate to wild-type enzyme and TrxR C138S stabilizes the enzymes in the FR conformation. Specific binding of the NADPH-analog, AADP(+), to the FR conformation resulted in dynamic fluorescence quenching in support of the multiple quenching sites model. Raising the temperature from 277K-323K resulted in a moderate shift to the FR conformation for TrxR C138S. High concentrations of the cosolvent glycerol triggered the domain rotation from the FO to the FR conformation.  相似文献   

16.
The fluorescence properties of various 8alpha-sulfur-linked flavinyl peptides and related flavin analogues were investigated as the pH solvent, temperature, and flavin concentration were varied. Substitution in the 8alpha position by a thioether-linked peptide brings about a marked quenching of fluorescence (up to 98% in water), a slight bathochromic shift and broadening of the fluorescence emission spectra, and a slight decrease in the fluorescence lifetimes. Oxidation of the thioether function to a sulfone partially releases this fluorescence quenching without further changes in the fluorescence emission spectra. The primary effect on the fluorescence intensity is due to an interaction between the nonbonding electrons of the thioether, the hydrogen-bonding, polar solvent, and the isoalloxazine ring. Dissolving these flavinyl peptides in nonaqueous solvents increases the fluorescence intensity as much as 20-fold. A secondary effect on flavinyl fluorescence can be attributed to a collisional quenching by the vicinal tyrosyl residue within tyrosine-containing flavinyl peptides. The fluorescence properties provide further confirmation of the identity of the synthetic and naturally obtained flavinyl peptides and of the interaction between the free-hydroxyl functions of the ribityl side chain and the thioether.  相似文献   

17.
The rotational motion of amphiphilic flavins in dipalmitoyl phospholipid bilayers was investigated with fluorescence anisotropy decay measurements. At temperatures between 10 and 50°C the rotation proved to be anisotropic, which indicated composite motion of both the aliphatic side-chain and the isoalloxazine moiety of the octadecyllumiflavin derivatives. Above the phase transition temperature (crystalline→liquid-crystalline state) the depolarization is complete within the average flavin fluorescence lifetime, implicating unrestricted motion and resulting in a non-ordered microenvironment. In the gel or crystalline state the flavin motion can best be characterized as a limited rotation or librational motion. The fluorescence decay of the flavins is heterogeneous at temperatures between 10 and 50°C, which is explained by assuming nanosecond relaxation of the polar phosphatidyl head-groups around the excited flavin. The lack of a significant cholesterol effect suggests that the isoalloxazine is located at the interphase of the bilayer and not in the hydrocarbon region. The microstructure is fluid-like, not in agreement with a preferred static localization of the flavins in the bilayer.  相似文献   

18.
This article describes novel data analysis of fluorescence lifetime-based protein kinase assays to identify and correct for compound interference in several practical cases. This ability, together with inherent advantages of fluorescence lifetime technology (FLT) as a homogeneous, antibody-free format independent of sample concentration, volume, excitation intensity, and geometry, makes fluorescence lifetime a practical alternative to the established “gold standards” of radiometric and mobility shift (Caliper) assays. The analysis is based on a photochemical model that sets constraints on the values of fluorescence lifetimes in the time responses of the assay. The addition of an exponential component with free floating lifetime to the constrained model, in which the lifetimes are constants predetermined from control measurements and the preexponential coefficients are “floating” parameters, allows the relative concentration of phosphorylated and nonphosphorylated substrates to be calculated even in the presence of compound fluorescence. The method is exemplified using both simulated data and experimental results measured from mixtures of dye-labeled phosphorylated and nonphosphorylated kinase substrates. A change of the fluorescence lifetime is achieved by the phosphorylated substrate-specific interaction with a bifunctional ligand, where one binding site interacts with the phosphate group and the other interacts with the dye.  相似文献   

19.
Time-resolved fluorescence and fluorescence anisotropy data surfaces of flavin adenine dinucleotide bound to lipoamide dehydrogenase from Azotobacter vinelandii in 80% glycerol have been obtained by variation of excitation energy and temperature between 203 and 303 K. The fluorescence kinetics of a deletion mutant lacking 14 COOH-terminal amino acids were compared with the wild-type enzyme to study a possible interaction of the COOH-terminal tail with the active site of the enzyme. The flavin adenine dinucleotide fluorescence in both proteins exhibits a bimodal lifetime distribution as recovered by the maximum entropy method of data analysis. The difference in standard enthalpy and entropy of associated conformational substates was retrieved from the fractional contributions of the two lifetime classes. Activation energies of thermal quenching were obtained that confirm that the isoalloxazines in the deletion mutant are solvent accessible in contrast to the wild-type enzyme. Red-edge spectroscopy in conjunction with variation of temperature provides the necessary experimental axes to interpret the fluorescence depolarization in terms of intersubunit energy transfer rather than reorientational dynamics of the flavins. The results can be explained by a compartmental model that describes the anisotropy decay of a binary, inhomogeneously broadened, homoenergy transfer system. By using this model in a global analysis of the fluorescence anisotropy decay surface, the distance between and relative orientation of the two isoalloxazine rings are elucidated. For the wild-type enzyme, this geometrical information is in agreement with crystallographic data of the A. vinelandii enzyme, whereas the mutual orientation of the subunits in the deletion mutant is slightly altered. In addition, the ambiguity in the direction of the emission transition moment in the isoalloxazine ring is solved. The anisotropy decay parameters also provide information on electronic and dipolar relaxational properties of the flavin active site. The local environment of the prosthetic groups in the deletion mutant of the A. vinelandii enzyme is highly inhomogeneous, and a transition from slow to rapid dipolar relaxation is observed over the measured temperature range. In the highly homogeneous active site of the wild-type enzyme, dipolar relaxation is slowed down beyond the time scale of fluorescence emission at any temperature studied. Our results are in favor of a COOH-terminal polypeptide interacting with the active site, thereby shielding the isoalloxazines from the solvent. This biological system forms a very appropriate tool to test the validity of photophysical models describing homoenergy transfer.  相似文献   

20.
This paper continues the studies of vesicle-bound flavins (‘anisotropic flavin chemistry’). It is possible to anchor the flavin nucleus in various modes within the lipid/water interface by means of long aliphatic chains and using different saturated lipids, thereby mimicking the specific binding of the coenzyme to the apoprotein in flavoproteins. Based on absorption spectroscopy and EPR spectroscopy studies we explored the rotational mobility and the microenvironment of membrane-bound amphiflavin radicals. N(5)-unsubstiluted amphiflavin radicals exhibit a similarly high disproportionation constant as known from isotropic flavin chemistry. However, reasonable stabilization of the radical was achieved by introduction of an alkyl group in position 5 in the reduced state prior to the one-electron oxidation. Adopting the fine structure of the corresponding EPR spectra as assay for the mobility of the semiquinone, we determined rotational relaxation times ranging from 60 ns in the crystalline state down to 10 or 15 ns in the liquid-crystalline state of the membrane. The solvatochromic effect shown by absorption spectra of the membrane-bound flavin radicals reflects a dielectric constant of the microenvironment of ? = 30–40, corresponding to the lipid/water interface region. The results obtained in this study are consistent with those obtained previously, from fluorescence analyses, supporting our former conclusions.  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