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 281 毫秒
1.
Thioredoxin reductase (TrxR) from Escherichia coli consists of two globular domains connected by a two-stranded beta sheet: an FAD domain and a pyridine nucleotide binding domain. The latter domain contains the redox-active disulfide composed of Cys 135 and Cys 138. TrxR is proposed to undergo a conformational change whereby the two domains rotate 66 degrees relative to each other (Waksman G, Krishna TSR, Williams CH Jr, Kuriyan J, 1994, J Mol Biol 236:800-816), placing either redox active disulfide (FO conformation) or the NADPH binding site (FR conformation) adjacent to the flavin. This domain rotation model was investigated by using a Cys 138 Ser active-site mutant. The flavin fluorescence of this mutant is only 7% that of wild-type TrxR, presumably due to the proximity of Ser 138 to the flavin in the FO conformation. Reaction of the remaining active-site thiol, Cys 135, with phenylmercuric acetate (PMA) causes a 9.5-fold increase in fluorescence. Titration of the PMA-treated mutant with the nonreducing NADP(H) analogue, 3-aminopyridine adenine dinucleotide phosphate (AADP+), results in significant quenching of the flavin fluorescence, which demonstrates binding adjacent to the FAD, as predicted for the FR conformation. Wild-type TrxR, with or without PMA treatment, shows similar quenching by AADP+, indicating that it exists mostly in the FR conformer. These findings, along with increased EndoGluC protease susceptibility of PMA-treated enzymes, agree with the model that the FO and FR conformations are in equilibrium. PMA treatment, because of steric limitations of the phenylmercuric adduct in the FO form, forces the equilibrium to the FR conformer, where AADP+ binding can cause fluorescence quenching and conformational restriction favors proteolytic susceptibility.  相似文献   

2.
Thioredoxin reductase (TrxR) from Escherichia coli, the mutant proteins E159Y and C138S, and the mutant protein C138S treated with phenylmercuric acetate were reconstituted with [U-(13)C(17),U-(15)N(4)]FAD and analysed, in their oxidized and reduced states, by (13)C-, (15)N- and (31)P-NMR spectroscopy. The enzymes studied showed very similar (31)P-NMR spectra in the oxidized state, consisting of two peaks at -9.8 and -11.5 p.p.m. In the reduced state, the two peaks merge into one apparent peak (at -9.8 p.p.m.). The data are compared with published (31)P-NMR data of enzymes closely related to TrxR. (13)C and (15)N-NMR chemical shifts of TrxR and the mutant proteins in the oxidized state provided information about the electronic structure of the protein-bound cofactor and its interactions with the apoproteins. Strong hydrogen bonds exist between protein-bound flavin and the apoproteins at C(2)O, C(4)O, N(1) and N(5). The N(10) atoms in the enzymes are slightly out of the molecular plane of the flavin. Of the ribityl carbon atoms C(10alpha,gamma,delta) are the most affected upon binding to the apoprotein and the large downfield shift of the C(10gamma) atom indicates strong hydrogen bonding with the apoprotein. The hydrogen bonding pattern observed is in excellent agreement with X-ray data, except for the N(1) and the N(3) atoms where a reversed situation was observed. Some chemical shifts observed in C138S deviate considerably from those of the other enzymes. From this it is concluded that C138S is in the FO conformation and the others are in the FR conformation, supporting published data. In the reduced state, strong hydrogen bonding interactions are observed between C(2)O and C(4)O and the apoprotein. As revealed by the (15)N chemical shifts and the N(5)H coupling constant the N(5) and the N(10) atom are highly sp(3) hybridized. The calculation of the endocyclic angles for the N(5) and the N(10) atoms shows the angles to be approximately 109 degrees, in perfect agreement with X-ray data showing that the flavin assumes a bent conformation along the N(10)/N(5) axis of the flavin. In contrast, the N(1) is highly sp(2) hybridized and is protonated, i.e. in the neutral state. Upon reduction of the enzymes, the (13)C chemical shifts of some atoms of the ribityl side chain undergo considerable changes also indicating conformational rearrangements of the side-chain interactions with the apoproteins. The chemical shifts between native TrxR and C138S are now rather similar and differ from those of the two other mutant proteins. This strongly indicates that the former enzymes are in the FO conformation and the other two are in the FR conformation. The data are discussed briefly in the context of published NMR data obtained with a variety of flavoproteins.  相似文献   

3.
Thioredoxin reductases (TrxRs) are flavin-containing dithioloxidoreductases that couple reduction equivalents from the soluble NAD(P)H pool to the soluble protein thioredoxin (Trx). Previous crystallographic studies of the Escherichia coli enzyme ( ecTrxR) have shown that low molecular weight TrxRs can adopt two distinct conformations: the first (FO) is required for the oxidation of the flavin cofactor and the generation of reduced Trx; the second (FR) is adopted for the reduction of the flavin by NAD(P)H. Here, protein electrochemistry has been used to interrogate the equilibrium between the oxidized and reduced conformations of the ecTrxR and a novel, low molecular weight TrxR from the thermophilic archaeon Thermoplasma acidophilum ( taTrxR) that is characterized structurally and biochemically in the accompanying paper [Hernandez et al. (2008) Biochemistry 47, 9728-9737]. A reversible electrochemical response is observed that reveals a dynamic behavior dependent upon the temperature of the experiment. At low temperatures (283 K) a broad, quasi-reversible electrochemical envelope is observed centered at a value of approximately -300 mV and displaying a peak width of over 150 mV. The voltammetric response sharpens dramatically as the temperature increases, becoming much more reversible (as determined by peak separation and peak width). The overall potential and shape of the voltammetric data indicate that the flavin (FAD/FADH 2) and disulfide/dithiol couples are very close in thermodynamic potentials, and the data are interpreted in terms of the model of two-state conformational change between flavin reducing (FR) and flavin oxidizing (FO) states, where the difference in potential for the flavin and disulfide cofactors must be within 40 mV of one another. In this model, the low temperature peak broadening is interpreted as an indication of a heterogeneous population of TrxR conformations that exist at low temperature; at higher temperatures, FO and FR conformers can rapidly interconvert, and voltammetry reports upon an average potential of the conformations.  相似文献   

4.
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   

5.
Catalysis by thioredoxin reductase (TrxR) from Escherichia coli requires alternation between two domain arrangements. One of these conformations has been observed by X-ray crystallography (Waksman G, Krishna TSR, Williams CH Jr, Kuriyan J, 1994, J Mol Biol 236:800-816). This form of TrxR, denoted FO, permits the reaction of enzyme-bound reduced FAD with a redox-active disulfide on TrxR. As part of an investigation of conformational changes and intermediates in catalysis by TrxR, an X-ray structure of the FO form of TrxR with both the FAD and active site disulfide reduced has been determined. Reduction after crystallization resulted in significant local conformation changes. The isoalloxazine ring of the FAD cofactor, which is essentially planar in the oxidized enzyme, assumes a 34 degree "butterfly" bend about the N(5)-N(10) axis in reduced TrxR. Theoretical calculations reported by others predict ring bending of 15-28 degrees for reduced isoalloxazines protonated at N(1). The large bending in reduced TrxR is attributed in part to steric interactions between the isoalloxazine ring and the sulfur of Cys138, formed by reduction of the active site disulfide, and is accompanied by changes in the positions and interactions of several of the ribityl side-chain atoms of FAD. The bending angle in reduced TrxR is larger than that for any flavoprotein in the Protein Data Bank. Distributions of bending angles in published oxidized and reduced flavoenzyme structures are different from those found in studies of free flavins, indicating that the protein environment has a significant effect on bending.  相似文献   

6.
Mutation of one of the cysteine residues in the redox active disulfide of thioredoxin reductase from Escherichia coli results in C135S with Cys138 remaining or C138S with Cys135 remaining. The expression system for the genes encoding thioredoxin reductase, wild-type enzyme, C135S, and C138S has been re-engineered to allow for greater yields of protein. Wild-type enzyme and C135S were found to be as previously reported, whereas discrepancies were detected in the characteristics of C138S. It was shown that the original C138S was a heterogeneous mixture containing C138S and wild-type enzyme and that enzyme obtained from the new expression system is the correct species. C138S obtained from the new expression system having 0.1% activity and 7% flavin fluorescence of wild-type enzyme was used in this study. Reductive titrations show that, as expected, only 1 mol of sodium dithionite/mol of FAD is required to reduce C138S. The remaining thiol in C135S and C138S has been reacted with 5,5'-dithiobis-(2-nitrobenzoic acid) to form mixed disulfides. The half time of the reaction was <5 s for Cys138 in C135S and approximately 300 s for Cys135 in C138S showing that Cys138 is much more reactive. The resulting mixed disulfides have been reacted with Cys32 in C35S mutant thioredoxin to form stable, covalent adducts C138S-C35S and C135S-C35S. The half times show that Cys138 is approximately fourfold more susceptible to attack by the nucleophile. These results suggest that Cys138 may be the thiol initiating dithiol-disulfide interchange between thioredoxin reductase and thioredoxin.  相似文献   

7.
The decay kinetics of the FAD-fluorescence in lipoamide dehydrogenase from pig heart have been reinvestigated using phase fluorometric methods and sophisticated laser pulse techniques. Both pulse and modulation methods lead to distinct heterogeneity in lifetimes. The two different techniques lead to good correspondence in the longer lifetime component of a biexponential decay model, whereas the more rapidly decaying component is distinctly shorter and has a larger amplitude using the phase technique with two available modulation frequencies (15 and 60 MHz). Lifetime measurements as a function of temperature and in the presence of D2O instead of H2O illustrate that the quenching of the FAD fluorescence in lipoamide. dehydrogenase is mainly dynamic in nature and that solvent comes into contact with the fluorophor. Mobility of the flavin itself, free and bound to the enzyme, has been measured by both differential polarized phase fluorometry and experimental fluorescence anisotropy decay after ps laser pulse excitation. By employing flavin models it has been shown that both techniques have ps time resolution. Measurements with the latter more direct method indicate a rapid subnanosecond motion of the FAD bound within the enzyme, only visible at temperatures lower than about 15°C, where the protein rotational diffusion is slowed down. The significance of rapid transient conformational fluctuations for catalysis is discussed with reference to recently developed insights reported in the literature.  相似文献   

8.
Chang SY  Chen YK  Wang AH  Liang PH 《Biochemistry》2003,42(49):14452-14459
Increasing evidence has shown that intrinsic disorder of proteins plays a key role in their biological functions. In the case of undecaprenyl pyrophosphate synthase (UPPs), which catalyzes the chain elongation of farnesyl pyrophosphate (FPP) to undecaprenyl pyrophosphate via eight consecutive condensation reactions with isopentenyl pyrophosphate, a highly flexible loop 72-83 was previously linked to protein conformational change required for catalysis [Chen, Y. H., Chen, A. P.-C., Chen, C. T., Wang, A. H.-J., and Liang, P. H., (2002) J. Biol. Chem. 277, 7369-7376]. The crystal structure and fluorescence studies suggested that the alpha3 helix connected to the loop moves toward the active site when the substrate is bound. To identify the active conformation and study the role of the loop for conformational change, the UPPs mutants with amino acids inserted into or deleted from the loop were examined. The inserted mutant with extra Ala residues fails to display the intrinsic fluorescence quenching upon FPP binding, and its crystal structure reveals only the open form. These phenomena appear to be different from the wild-type enzyme in which open and closed conformers were observed and suggest that the extended loop fails to pull the alpha3 helix and/or the extra amino acids in the loop cause steric hindrance on the alpha3 helix movement. The loop-shortening mutants with deletion of V82 and S83 or S72 also adopt an open conformation with the loop stretched, although they show decreased intrinsic fluorescence with FPP bound, similar to that seen in the wild-type enzyme. We conclude that the closed conformation is apparently the active conformation. Change of the length of the loop 72-83 impairs the ability of conformational change and causes remarkably lower activity of UPPs.  相似文献   

9.
Trimethylamine dehydrogenase (TMADH) is an iron-sulfur flavoprotein that catalyzes the oxidative demethylation of trimethylamine to form dimethylamine and formaldehyde. It contains a unique flavin, in the form of a 6-S-cysteinyl FMN, which is bent by approximately 25 degrees along the N5-N10 axis of the flavin isoalloxazine ring. This unusual conformation is thought to modulate the properties of the flavin to facilitate catalysis, and has been postulated to be the result of covalent linkage to Cys-30 at the flavin C6 atom. We report here the crystal structures of recombinant wild-type and the C30A mutant TMADH enzymes, both determined at 2.2 A resolution. Combined crystallographic and NMR studies reveal the presence of inorganic phosphate in the FMN binding site in the deflavo fraction of both recombinant wild-type and C30A proteins. The presence of tightly bound inorganic phosphate in the recombinant enzymes explains the inability to reconstitute the deflavo forms of the recombinant wild-type and C30A enzymes that are generated in vivo. The active site structure and flavin conformation in C30A TMADH are identical to those in recombinant and native TMADH, thus revealing that, contrary to expectation, the 6-S-cysteinyl FMN link is not responsible for the 25 degrees butterfly bending along the N5-N10 axis of the flavin in TMADH. Computational quantum chemistry studies strongly support the proposed role of the butterfly bend in modulating the redox properties of the flavin. Solution studies reveal major differences in the kinetic behavior of the wild-type and C30A proteins. Computational studies reveal a hitherto, unrecognized, contribution made by the S(gamma) atom of Cys-30 to substrate binding, and a role for Cys-30 in the optimal geometrical alignment of substrate with the 6-S-cysteinyl FMN in the enzyme active site.  相似文献   

10.
Using molecular dynamics simulations and steady‐state fluorescence spectroscopy, we have identified a conformational change in the active site of a thermophilic flavoenzyme, NADH oxidase from Thermus thermophilus HB8 (NOX). The enzyme's far‐UV circular dichroism spectrum, intrinsic tryptophan fluorescence, and apparent molecular weight measured by dynamic light scattering varied little between 25 and 75°C. However, the fluorescence of the tightly bound FAD cofactor increased approximately fourfold over this temperature range. This effect appears not to be due to aggregation, unfolding, cofactor dissociation, or changes in quaternary structure. We therefore attribute the change in flavin fluorescence to a temperature‐dependent conformational change involving the NOX active site. Molecular dynamics simulations and the effects of mutating aromatic residues near the flavin suggest that the change in fluorescence results from a decrease in quenching by electron transfer from tyrosine 137 to the flavin. Proteins 2012. © 2011 Wiley Periodicals, Inc.  相似文献   

11.
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   

12.
Chen H  Zhang G  Zhang Y  Dong Y  Yang K 《Biochemistry》2000,39(40):12140-12148
Prochymosin (chymosin) contains three disulfide bonds: Cys45-Cys50, Cys206-Cys210, and Cys250-Cys283. We have demonstrated that Cys250-Cys283 is indispensable for correct refolding of prochymosin, whereas Cys45-Cys50 is dispensable but has some contribution to the stability and substrate specificity of the enzyme. Here, we report the results about the functions of Cys206-Cys210 by site-directed mutagenesis studies. In a glutathione redox system C206A/C210A mutant exhibited oxidative refolding kinetics and efficiency ( approximately 40% reactivation) similar to those of the wild-type prochymosin, indicating that Cys206-Cys210 is also dispensable for refolding. However, C206S/C210S and single-site mutants (C210A, C210S, and C206A) showed only about 3 and 0-0.4% reactivation, respectively. This is quite different from the Cys45-Cys50 deficient mutants (C45A, C50A, C45A/C50A, C45D, C50S, C45D/C50S, C45A/C50S), which have comparable refolding efficiencies, implying that the substituents at position 206 and 210 play more important role in determining correct refolding than those at position 45 and 50. Urea-induced denaturation and fluorescence quenching studies indicated that the prochymosin mutants C206A/C210A and C206S/C210S were 2.1 and 4.8 kJ/mol less stable than prochymosin and some tryptophan residue in the mutated molecules was less exposed. However, the wild-type and mutant prochymosins shared similar far-UV CD and fluorescence emission spectra and similar specific potential activity, suggesting that the overall conformation was maintained after mutation. Activity assay and kinetic analysis revealed that mutation did not change the specific milk-clotting activity significantly but resulted in an increase in K(m) and k(cat) toward a hexapeptide substrate. On the basis of the above-mentioned perturbance of tryptophanyl microenvironment and the three-dimensional structure of chymosin, we proposed that deletion of Cys206-Cys210 may induce a propagated conformational change, resulting in a perturbance of the local conformation around active-site cleft and in turn, an alteration of the substrate specificity.  相似文献   

13.
14.
A study was made of the processes associated with the quenching of 4″-dimethylaminochalcone (DMAC) fluorescence by proton-donor solvent (1-butanol). The kinetics of deactivation of the DMAC excited state was assessed by transient absorption spectra with a time resolution about 50 fs and by fluorescence decay with ~30-ps resolution. The following sequence of events could thus be envisaged: (i) the DMAC molecule in the ground state (prior to excitation) makes a hydrogen bond with an alcohol molecule; (ii) absorption of a light quantum causes a corresponding increase of the DMAC dipole moment; the H-bond is retained; (iii) the solvation shell formed by alcohol dipoles is reorganized in response to the raise of the DMAC dipole moment, with an energy expenditure about 24 kJ/mol and a time constant about 40 ps; the initial H-bond is still retained; (iv) processes leading to fluorescence quenching occur with an effective time constant of nearly 200 ps. Since quenching is far slower than solvate rearrangement, one can suppose that it is not a direct consequence of shell relaxation or prior H-bonding. Thus, DMAC fluorescence quenching may involve different processes observed with other aromatic molecules: H-bond rearrangement from a nonquenching to a more ‘efficient’ conformation, charge transfer between the excited molecule and alcohol, or solvent-induced out-of-plane twist of the DMAC amino group.  相似文献   

15.
Secundo F  Russo C  Giordano A  Carrea G  Rossi M  Raia CA 《Biochemistry》2005,44(33):11040-11048
A combination of hydrogen/deuterium exchange, fluorescence quenching, and kinetic studies was used to acquire experimental evidence for the crystallographically hypothesized increase in local flexibility which occurs in thermophilic NAD(+)-dependent Sulfolobus solfataricus alcohol dehydrogenase (SsADH) upon substitution Asn249Tyr. The substitution, located at the adenine-binding site, proved to decrease the affinity for both coenzyme and substrate, rendering the mutant enzyme 6-fold more active when compared to the wild-type enzyme [Esposito et al. (2003) FEBS Lett. 539, 14-18]. The amide H/D exchange data show that the wild-type and mutant enzymes have similar global flexibility at 22 and 60 degrees C. However, the temperature dependence of the Stern-Volmer constant determined by acrylamide quenching shows that the increase in temperature affects the local flexibility differently, since the K(SV) increment is significantly higher for the wild-type than for the mutant enzyme over the range 18-45 degrees C. Interestingly, the corresponding van't Hoff plot (log K(SV) vs 1/T) proves nonlinear for the apo and holo wild-type and apo mutant enzymes, with a break at approximately 45 degrees C in all three cases due to a conformational change affecting the tryptophan microenvironment experienced by the quencher molecules. The Arrhenius and van't Hoff plots derived from the k(cat) and K(M) thermodependence measured with cyclohexanol and NAD(+) at different temperatures display an abrupt change of slope at 45-50 degrees C. This proves more pronounced in the case of the mutant enzyme compared to the wild-type enzyme due to a conformational change in the structure rather than to an overlapping of two or more rate-limiting reaction steps with different temperature dependencies of their rate constants. Three-dimensional analysis indicates that the observed conformational change induced by temperature is associated with the flexible loops directly involved in the substrate and coenzyme binding.  相似文献   

16.
Sulfolobus solfataricus 5'-deoxy-5'-melthylthioadenosine phosphorylase II (SsMTAPII), is a hyperthermophilic hexameric protein with two intrasubunit disulfide bonds (C138-C205 and C200-C262) and a CXC motif (C259-C261). To get information on the role played by these covalent links in stability and folding, the conformational stability of SsMTAPII and C262S and C259S/C261S mutants was studied by thermal and guanidinium chloride (GdmCl)-induced unfolding and analyzed by fluorescence spectroscopy, circular dichroism, and SDS-PAGE. No thermal unfolding transition of SsMTAPII can be obtained under nonreducing conditions, while in the presence of the reducing agent Tris-(2-carboxyethyl) phosphine (TCEP), a Tm of 100°C can be measured demonstrating the involvement of disulfide bridges in enzyme thermostability. Different from the wild-type, C262S and C259S/C261S show complete thermal denaturation curves with sigmoidal transitions centered at 102°C and 99°C respectively. Under reducing conditions these values decrease by 4°C and 8°C respectively, highlighting the important role exerted by the CXC disulfide on enzyme thermostability. The contribution of disulfide bonds to the conformational stability of SsMTAPII was further assessed by GdmCl-induced unfolding experiments carried out under reducing and nonreducing conditions. Thermal unfolding was found to be reversible if the protein was heated in the presence of TCEP up to 90°C but irreversible above this temperature because of aggregation. In analogy, only chemical unfolding carried out in the presence of reducing agents resulted in a reversible process suggesting that disulfide bonds play a role in enzyme denaturation. Thermal and chemical unfolding of SsMTAPII occur with dissociation of the native hexameric state into denatured monomers, as indicated by SDS-PAGE.  相似文献   

17.
Frederick KK  Palfey BA 《Biochemistry》2005,44(40):13304-13314
p-Hydroxybenzoate hydroxylase (PHBH) is an FAD-dependent monooxygenase that catalyzes the hydroxylation of p-hydroxybenzoate (pOHB) to 3,4-dihydroxybenzoate in an NADPH-dependent reaction. Two structural features are coupled to control the reactivity of PHBH with NADPH: a proton-transfer network that allows protons to be passed between the sequestered active site and solvent and a flavin that adopts two positions: "in", where the flavin is near pOHB, and "out", where the flavin is near NADPH. PHBH uses the proton-transfer network to test for the presence of a suitable aromatic substrate before allowing the flavin to adopt the NADPH-accessible conformation. In this work, kinetic analysis of the His72Asn mutant, with a disrupted proton-transfer network, showed that flavin movement could occur in the presence or absence of NADPH but that NADPH stimulated movement to the reactive conformation required for hydride transfer. Substrate and solvent isotope effects on the transient kinetics of reduction of the His72Asn mutant showed that proton transfer was linked to flavin movement and that the conformational change occurred in a step separate from that of hydride transfer. Proton transfers during the reductive half-reaction were observed directly in the wild-type enzyme by performing experiments in the presence of a fluorescent pH-indicator dye in unbuffered solutions. NADPH binding caused rapid proton release from the enzyme, followed by proton uptake after flavin reduction. Solvent and substrate kinetic isotope effects showed that proton-coupled flavin movement and reduction also occurred in different steps in wild-type PHBH. These results allow a detailed kinetic scheme to be proposed for the reductive half-reaction of the wild-type enzyme. Three kinetic models considered for substrate-induced isomerization are analyzed in the Appendix.  相似文献   

18.
Although dimerization appears to be a common property of G-protein-coupled receptors (GPCRs), it remains unclear whether a GPCR dimer binds one or two molecules of ligand and whether ligand binding results in activation of one or two G-proteins when measured using functional assays in intact living cells. Previously, we demonstrated that serotonin 5-hydroxytryptamine2C (5-HT2C) receptors form homodimers (Herrick-Davis, K., Grinde, E., and Mazurkiewicz, J. (2004) Biochemistry 43, 13963-13971). In the present study, an inactive 5-HT(2C) receptor was created and coexpressed with wild-type 5-HT2C receptors to determine whether dimerization regulates receptor function and to determine the ligand/dimer/G-protein stoichiometry in living cells. Mutagenesis of Ser138 to Arg (S138R) produced a 5-HT2C receptor incapable of binding ligand or stimulating inositol phosphate (IP) signaling. Confocal fluorescence imaging revealed plasma membrane expression of yellow fluorescent protein-tagged S138R receptors. Expression of wild-type 5-HT2C receptors in an S138R-expressing stable cell line had no effect on ligand binding to wild-type 5-HT2C receptors, but inhibited basal and 5-HT-stimulated IP signaling as well as constitutive and 5-HT-stimulated endocytosis of wild-type 5-HT2C receptors. M1 muscarinic receptor activation of IP production was normal in the S138R-expressing cells. Heterodimerization of S138R with wild-type 5-HT2C receptors was visualized in living cells using confocal fluorescence resonance energy transfer (FRET). FRET was dependent on the donor/acceptor ratio and independent of the receptor expression level. Therefore, inactive 5-HT2C receptors inhibit wild-type 5-HT2C receptor function by forming nonfunctional heterodimers expressed on the plasma membrane. These results are consistent with a model in which one GPCR dimer binds two molecules of ligand and one G-protein and indicate that dimerization is essential for 5-HT receptor function.  相似文献   

19.
We have been investigated the relation between activation of "neutral" and "acidic" chymotrypsin-like (ChT-L) activity and conformational changes in the 20S proteasome complex from the rat natural killer (NK) cells induced by SDS, mono- and divalent cations. The conformational changes were monitored by tryptophan fluorescence and light scattering. It was revealed that the changes in the maximum position and contribution of the short-wavelength spectral component correlated with the alteration of ChT-L activity of the proteasome. Statistical analysis was applied to assign the fluorescence components with tryptophan residues based on the classification of calculated structural parameters of the environment of tryptophan fluorophores in protein. It was proposed that the emission of W13 from [Formula: see text] -subunit located near the cluster of highly conserved proteasome residues is mostly sensitive to the activation of the enzyme. We concluded that the expression of maximal ChT-L activity of 20S proteasome is associated with the conformational changes occurs in this cluster that lead to the proteasome open conformation, allowing substrate access into the proteolytic chamber.  相似文献   

20.
1. The NADP-dependent glutamate dehydrogenase of Neurospora crassa undergoes slow reversible structural transitions, with half-times in the order of a few minutes, between active and inactive states. The inactive state of the enzyme, which predominates at pH values below 7.0, has an intrinsic tryptophan fluorescence 25% lower than that of the active state, which predominates at pH values above 7.6. The inactive state can be activated either by an increase in pH or by addition of activators such as succinate. 2. The kinetics of the slow transitions that follow activating and inactivating rapid changes in conditions have been monitored by measurements of protein fluorescence. The results show that the slow reversible conformational change detected by the change in fluorescence is the rate-limiting process for enzyme activation and inactivation. 3. In both directions this conformational change follows apparent first-order kinetics and the rate constant is independent of protein concentration. These kinetics and published measurements of molecular weight are indicative of an isomerization process. 4. In both directions the changes show a large energy of activation and a large positive entropy of activation, consistent with a considerable disturbance of conformation in the transition state. 5. Comparisons of the fluorescence emission spectra of the active and inactive states indicate that the difference in fluorescence is produced by quenching, possibly intramolecular, in the inactive conformation. Iodide ions cause similar quenching. 6. In some mutationally altered forms of the enzyme comparable but modified conformational changes can be followed by protein fluorescence.  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