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.
Human methionine synthase reductase (MSR) catalyzes the NADPH-dependent reductive methylation of methionine synthase. MSR is 78 kDa flavoprotein belonging to a family of diflavin reductases, with cytochrome P450 reductase (CPR) as the prototype. MSR and its individual flavin-binding domains were cloned as GST-tagged fusion proteins for expression and purification from Escherichia coli. The isolated flavin domains of MSR retain UV-visible and secondary structural properties indicative of correctly folded flavoproteins. Anaerobic redox titrations on the individual domains assisted in assignment of the midpoint potentials for the high- and low-potential flavin. For the isolated FMN domain, the midpoint potentials for the oxidized/semiquinone (ox/sq) couple and semiquinone/hydroquinone (sq/hq) couple are -112 and -221 mV, respectively, at pH 7.0 and 25 degrees C. The corresponding couples in the isolated FAD domain are -222 mV (ox/sq) and -288 mV (sq/hq). Both flavins form blue neutral semiquinone species characterized by broad absorption peaks in the long-wavelength region during anaerobic titration with sodium dithionite. In full-length MSR, the values of the FMN couples are -109 mV (ox/sq) and -227 mV (sq/hq), and the corresponding couple values for FAD are -254 mV (ox/sq) and -291 mV (sq/hq). Separation of the MSR flavins does not perturb their thermodynamic properties, as midpoint potentials for all four couples are similar in isolated domains and in full-length MSR. The redox properties of MSR are discussed in relation to other members of the diflavin oxidoreductase family and the mechanism of electron transfer.  相似文献   

2.
Human short-chain acyl-CoA dehydrogenase (hSCAD) catalyzes the first matrix step in the mitochondrial beta-oxidation cycle for substrates with four and six carbons. Previous studies have shown that the act of substrate/product binding induces a large enzyme potential shift in acyl-CoA dehydrogenases. The objective of this work was to examine the thermodynamic regulation of this process through direct characterization of the electrochemical properties of hSCAD using spectroelectrochemical methodology. A large amount of substrate activation was observed in the enzymatic reaction of hSCAD (+33 mV), the greatest magnitude measured in any acyl-CoA dehydrogenase to date. To examine the role of the substrate as well as the product in electron transfer by hSCAD, a catalytic base mutation (E368Q) was constructed. The E368Q mutation inactivates the reductive and oxidative pathways such that the individual effects of substrate and product binding on the redox potential can be investigated. Optimal substrate (butyryl-CoA) was seen to shift the flavin redox potential slightly more positive (+38 mV) than did optimal product (crotonyl-CoA) (+31 mV), a finding opposite of that observed in another short-chain enzyme, bacterial SCAD. These results indicate that substrate redox activation occurs in hSCAD leading to a large enzyme midpoint potential shift. Substrate binding in hSCAD appears to make a larger contribution than does product to thermodynamic modulation.  相似文献   

3.
L H Bradley  R P Swenson 《Biochemistry》1999,38(38):12377-12386
The midpoint potentials for both redox couples of the noncovalently bound flavin mononucleotide (FMN) cofactor in the flavodoxin are known to be pH dependent. While the pH dependency for the oxidized-semiquinone (ox/sq) couple is consistent with the formation of the blue neutral form of the flavin semiquinone, that of the semiquinone-hydroquinone (sq/hq) couple is more enigmatic. The apparent pK(a) of 6.7 for this couple in the flavodoxin from Clostridium beijerinckii has been attributed to the ionization of the FMN(HQ); however, nuclear magnetic resonance data strongly suggest the FMN(HQ) remains anionic over the entire pH range testable. As an alternative explanation, a specific glutamate residue (Glu59 in this flavodoxin), which is hydrogen-bonded to N(3)H of the FMN, has been postulated to be the primary redox-linked proton acceptor responsible for the pH effect in some flavodoxins. This model was directly tested in this study by permanently neutralizing Glu59 by its replacement with glutamine. This conservative substitution resulted in an increase of 86 mV (at pH 7) in midpoint potential of the sq/hq couple; however, the pH dependency of this couple was not altered. Thus, the redox-linked protonation of Glu59 clearly cannot be responsible for this effect as proposed. The pH dependency of the ox/sq couple was also similar to wild type, but the midpoint potential has decreased by 65 mV (pH 7). The K(d) values for the oxidized, semiquinone, and hydroquinone complexes increased by 43-, 590-, and 20-fold, respectively, relative to the wild type. Thus, the Glu59 to glutamine substitution substantially effects the stability of the semiquinone but, on a relative basis, slightly favors the formation of the hydroquinone. On the basis of (1)H-(15)N HSQC nuclear magnetic resonance spectroscopic studies, the increased temperature coefficients for the protons on N(3) and N(5) of the reduced FMN in E59Q suggest that the hydrogen-bonding interactions at these positions are significantly weakened in this mutant. The increase for N(5)H correlates with the reduced stability of the FMN(SQ) and the more negative midpoint potential for the ox/sq couple. On the basis of the X-ray structure, an "anchoring" role is proposed for the side chain carboxylate of Glu59 that stabilizes the structure of the 50's loop in such a way so as to promote the crucial hydrogen-bonding interaction that stabilizes the flavin semiquinone, contributing to the low potential of this flavodoxin.  相似文献   

4.
Oxidation-reduction potentials of butyryl-CoA dehydrogenase   总被引:1,自引:0,他引:1  
In order to obtain butyryl-CoA dehydrogenase from Megasphaera elsdenii in pure enough form to perform redox studies, the existing purification procedures first had to be modified and clarified [Engel, P. (1981) Methods Enzymol. 71, 359-366]. These modifications are described, and the previously unpublished spectral properties of the electrophoretically pure CoA-free butyryl-CoA dehydrogenase are presented. In our spectral reductive titration of pure enzyme, we show that although blue neutral flavin radical is stabilized in nonquantitative amounts in dithionite titrations (19%) or in electrochemical reductions mediated by methylviologen (5%), it is not thermodynamically stabilized; therefore, only a midpoint potential for butyryl-CoA dehydrogenase is obtained. The electron-transfer behavior from pH 5.5 to pH 7.0 indicates reversible two-electron transfer accompanied by one proton: EFlox + 2e- + H+ = EFlredH- Em7 = -0.079 V vs. SHE where EFlox is oxidized butyryl-CoA dehydrogenase, EFlredH- is two electron reduced enzyme, and Em7 is the midpoint potential at pH 7.0 at 25 degrees C. Redox data and activity data both indicate that the enzyme loses activity rapidly at pH values above 7.0. The Em7 of the butyryl-CoA dehydrogenase is 40 mV positive of the Em7 of the butyryl-CoA/crotonyl-CoA couple [Gustafson, W. G., Feinberg, B. A., & McFarland, J. T. (1986) J. Biol. Chem. 261, 7733-7741]. Binding of substrate analogue acetoacetyl-CoA caused the potential of butyryl-CoA dehydrogenase to shift 100 mV negative of the free enzyme. The negative shift in potential makes electron transfer from enzyme to substrate more probable, which is consistent with the direction of electron transfer in the bacterial system.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The conserved sequence motif "RxY(T)(S)xx(S)(N)" coordinates flavin binding in NADH:cytochrome b(5) reductase (cb(5)r) and other members of the flavin transhydrogenase superfamily of oxidoreductases. To investigate the roles of Y93, the third and only aromatic residue of the "RxY(T)(S)xx(S)(N)" motif, that stacks against the si-face of the flavin isoalloxazine ring, and P92, the second residue in the motif that is also in close proximity to the FAD moiety, a series of rat cb(5)r variants were produced with substitutions at either P92 or Y93, respectively. The proline mutants P92A, G, and S together with the tyrosine mutants Y93A, D, F, H, S, and W were recombinantly expressed in E. coli and purified to homogeneity. Each mutant protein was found to bind FAD in a 1:1 cofactor:protein stoichiometry while UV CD spectra suggested similar secondary structure organization among all nine variants. The tyrosine variants Y93A, D, F, H, and S exhibited varying degrees of blue-shift in the flavin visible absorption maxima while visible CD spectra of the Y93A, D, H, S, and W mutants exhibited similar blue-shifted maxima together with changes in absorption intensity. Intrinsic flavin fluorescence was quenched in the wild type, P92S and A, and Y93H and W mutants while Y93A, D, F, and S mutants exhibited increased fluorescence when compared to free FAD. The tyrosine variants Y93A, D, F, and S also exhibited greater thermolability of FAD binding. The specificity constant (k(cat)/K(m)(NADH)) for NADH:FR activity decreased in the order wild type > P92S > P92A > P92G > Y93F > Y93S > Y93A > Y93D > Y93H > Y93W with the Y93W variant retaining only 0.5% of wild-type efficiency. Both K(s)(H4NAD) and K(s)(NAD+) values suggested that Y93A, F, and W mutants had compromised NADH and NAD(+) binding. Thermodynamic measurements of the midpoint potential (E degrees ', n = 2) of the FAD/FADH(2) redox couple revealed that the potentials of the Y93A and S variants were approximately 30 mV more positive than that of wild-type cb(5)r (E degrees ' = -268 mV) while that of Y93H was approximately 30 mV more negative. These results indicate that neither P92 nor Y93 are critical for flavin incorporation in cb(5)r and that an aromatic side chain is not essential at position 93, but they demonstrate that Y93 forms contacts with the FAD that effectively modulate the spectroscopic, catalytic, and thermodynamic properties of the bound cofactor.  相似文献   

6.
The long-chain flavodoxins, with 169-176 residues, display oxidation-reduction potentials at pH 7 that vary from -50 to -260 mV for the oxidized/semiquinone (ox/sq) equilibrium and are -400 mV or lower for the semiquinone/hydroquinone (sq/hq) equilibrium. To examine the effects of protein interactions and conformation changes on FMN potentials in the long-chain flavodoxin from Anacystis nidulans (Synechococcus PCC 7942), we have determined crystal structures for the semiquinone and hydroquinone forms of the wild-type protein and for the mutant Asn58Gly, and have measured redox potentials and FMN association constants. A peptide near the flavin ring, Asn58-Val59, reorients when the FMN is reduced to the semiquinone form and adopts a conformation ("O-up") in which O 58 hydrogen bonds to the flavin N(5)H; this rearrangement is analogous to changes observed in the flavodoxins from Clostridium beijerinckii and Desulfovibrio vulgaris. On further reduction to the hydroquinone state, the Asn58-Val59 peptide in crystalline wild-type A. nidulans flavodoxin rotates away from the flavin to the "O-down" position characteristic of the oxidized structure. This reversion to the conformation found in the oxidized state is unusual and has not been observed in other flavodoxins. The Asn58Gly mutation, at the site which undergoes conformation changes when FMN is reduced, was expected to stabilize the O-up conformation found in the semiquinone oxidation state. This mutation raises the ox/sq potential by 46 mV to -175 mV and lowers the sq/hq potential by 26 mV to -468 mV. In the hydroquinone form of the Asn58Gly mutant the C-O 58 remains up and hydrogen bonded to N(5)H, as in the fully reduced flavodoxins from C. beijerinckii and D. vulgaris. The redox and structural properties of A. nidulans flavodoxin and the Asn58Gly mutant confirm the importance of interactions made by N(5) or N(5)H in determining potentials, and are consistent with earlier conclusions that conformational energies contribute to the observed potentials.The mutations Asp90Asn and Asp100Asn were designed to probe the effects of electrostatic interactions on the potentials of protein-bound flavin. Replacement of acidic by neutral residues at positions 90 and 100 does not perturb the structure, but has a substantial effect on the sq/hq equilibrium. This potential is increased by 25-41 mV, showing that electrostatic interaction between acidic residues and the flavin decreases the potential for conversion of the neutral semiquinone to the anionic hydroquinone. The potentials and the effects of mutations in A. nidulans flavodoxin are rationalized using a thermodynamic scheme developed for C. beijerinckii flavodoxin.  相似文献   

7.
Until now, workers in the field of fatty acid metabolism have suggested that the substrates are isopotential with the enzymes and that the reactions are forced to completion by the formation of charge-transfer complexes [Gustafson, W. G., Feinberg, B. A., & McFarland, J. T. (1986) J. Biol. Chem. 261, 7733-7741]. To date, no experimental evidence for this hypothesis exists. The work presented here shows that the butyryl-CoA/crotonyl-CoA couple is not isopotential with the enzymes with which it interacts. The potential of the butyryl-CoA/crotonyl-CoA couple (E ' = -0.013 V) is significantly more positive than the potential of either of the enzymes with which it interacts, bacterial butyryl-CoA dehydrogenase (E ' = -0.079 V) and mammalian general acyl-CoA dehydrogenase (E ' = 0.133 V). These data imply that the regulation of enzyme potential is essential for any electron transfer from substrate to enzyme to occur in mammalian or bacterial systems. In support of this assertion, a significant shift in potential for bacterial butyryl-CoA dehydrogenase (an analogue of the mammalian enzyme) in the presence of butyryl-CoA and crotonyl-CoA is reported. The potential is shifted positive by 60 mV. Larger potential shifts will undoubtedly be observed with the mammalian enzyme, which would be consistent with the catalytic direction of electron transfer.  相似文献   

8.
We have determined reduction potentials for porcine mitochondrial general fatty acyl-CoA dehydrogenase (GAD) and electron transfer flavoprotein (ETF) using an anaerobic spectroelectrochemical titration method. Computer simulation techniques were used to analyze the absorbance data. Nernst plots of the simulated data gave E'0, 7.1, quinone/semiquinone = -0.014 V and E'0, 7.1, semiquinone/hydroquinone = -0.036 V for ETF and E'0, 7.1, quinone/semiquinone = -0.155 V and E'0, 7.1, semiquinone/hydroquinone = -0.122 V for GAD. Using these techniques we have also determined a conditional reduction potential of -0.156 V for the chromophore producing fatty acyl-CoA substrate beta-2-furylpropionyl-CoA. From this value and our previous determination of the equilibrium constant for the transhydrogenation reaction between beta-2-furylpropionyl-CoA and the oxidized substrate crotonyl-CoA (Keq = 10.4), we have determined a reduction potential of -0.126 V for the butyryl-CoA/crotonyl-CoA couple. In light of the structural similarity between butyryl-CoA and octanoyl-CoA, the optimal substrate for GAD, the reduction potential for octanoyl-CoA should be similar to that for butyryl-CoA; i.e. fatty acyl-CoA substrates and GAD are essentially isopotential. The ability of octanoyl-CoA to reduce GAD quantitatively (Keq = 9.0) poses a dilemma in light of the nearly equal reduction potentials. We postulate that the stable charge-transfer complex formed between enzyme and optimal product is significantly lower in energy than enzyme and product and thus is responsible for pulling the reaction toward completion.  相似文献   

9.
Flavoproteins can dramatically adjust the thermodynamics and kinetics of electron transfer at their flavin cofactor. A versatile regulatory tool is proton transfer. Here, we demonstrate the significance of proton-coupled electron transfer to redox tuning and semiquinone (sq) stability in photolyases (PLs) and cryptochromes (CRYs). These light-responsive proteins share homologous overall architectures and FAD-binding pockets, yet they have evolved divergent functions that include DNA repair, photomorphogenesis, regulation of circadian rhythm, and magnetoreception. We report the first measurement of both FAD redox potentials for cyclobutane pyrimidine dimer PL (CPD-PL, Anacystis nidulans). These values, E(1)(hq/sq) = -140 mV and E(2)(sq/ox) = -219 mV, where hq is FAD hydroquinone and ox is oxidized FAD, establish that the sq is not thermodynamically stabilized (ΔE = E(2) - E(1) = -79 mV). Results with N386D CPD-PL support our earlier hypothesis of a kinetic barrier to sq oxidation associated with proton transfer. Both E(1) and E(2) are upshifted by ~ 100 mV in this mutant; replacing the N5-proximal Asn with Asp decreases the driving force for sq oxidation. However, this Asp alleviates the kinetic barrier, presumably by acting as a proton shuttle, because the sq in N386D CPD-PL oxidizes orders of magnitude more rapidly than wild type. These data clearly reveal, as suggested for plant CRYs, that an N5-proximal Asp can switch on proton transfer and modulate sq reactivity. However, the effect is context-dependent. More generally, we propose that PLs and CRYs tune the properties of their N5-proximal residue to adjust the extent of proton transfer, H-bonding patterns, and changes in protein conformation associated with electron transfer at the flavin.  相似文献   

10.
Ghanem M  Gadda G 《Biochemistry》2006,45(10):3437-3447
A protein positive charge near the flavin N(1) locus is a distinguishing feature of most flavoprotein oxidases, with mechanistic implications for the modulation of flavin reactivity. A recent study showed that in the active site of choline oxidase the protein positive charge is provided by His(466). Here, we have reversed the charge by substitution with aspartate (CHO-H466D) and, for the first time, characterized a flavoprotein oxidase with a negative charge near the flavin N(1) locus. CHO-H466D formed a stable complex with choline but lost the ability to oxidize the substrate. In contrast to the wild-type enzyme, which binds FAD covalently in a 1:1 ratio, CHO-H466D contained approximately 0.3 FAD per protein, of which 75% was not covalently bound to the enzyme. Anaerobic reduction of CHO-H466D resulted in the formation of a neutral hydroquinone, with no stabilization of the flavin semiquinone; in contrast, the anionic semiquinone and hydroquinone species were observed with the wild type and a H466A variant of the enzyme. The midpoint reduction potential for the oxidized-reduced couple in CHO-H466D was approximately 160 mV lower than that of the wild-type enzyme. Finally, CHO-H466D lost the ability to form complexes with glycine betaine or sulfite. Thus, with a reversal of the protein charge near the FAD N(1) locus, choline oxidase lost the ability to stabilize negative charges in the active site, irrespective of whether they develop on the flavin or are borne on ligands, resulting in defective flavinylation of the protein, the decreased electrophilicity of the flavin, and the consequent loss of catalytic activity.  相似文献   

11.
4-Hydroxybutyryl-coenzyme A (CoA) dehydratase (4HBD) from Clostridium aminobutyricum catalyzes the reversible dehydration of 4-hydroxybutyryl-CoA to crotonyl-CoA and the irreversible isomerization of vinylacetyl-CoA to crotonyl-CoA. 4HBD is an oxygen-sensitive homotetrameric enzyme with one [4Fe-4S]2+ cluster and one flavin adenine dinucleotide (FAD) in each subunit. Upon the addition of crotonyl-CoA or the analogues butyryl-CoA, acetyl-CoA, and CoA, UV-visible light and electron paramagnetic resonance (EPR) spectroscopy revealed an internal one-electron transfer to FAD and the [4Fe-4S]2+ cluster prior to hydration. We describe an active recombinant 4HBD and variants produced in Escherichia coli. The variants of the cluster ligands (H292C [histidine at position 292 is replaced by cysteine], H292E, C99A, C103A, and C299A) had no measurable dehydratase activity and were composed of monomers, dimers, and tetramers. Variants of other potential catalytic residues were composed only of tetramers and exhibited either no measurable (E257Q, E455Q, and Y296W) hydratase activity or <1% (Y296F and T190V) dehydratase activity. The E455Q variant but not the Y296F or E257Q variant displayed the same spectral changes as the wild-type enzyme after the addition of crotonyl-CoA but at a much lower rate. The results suggest that upon the addition of a substrate, Y296 is deprotonated by E455 and reduces FAD to FADH·, aided by protonation from E257 via T190. In contrast to FADH·, the tyrosyl radical could not be detected by EPR spectroscopy. FADH· appears to initiate the radical dehydration via an allylic ketyl radical that was proposed 19 years ago. The mode of radical generation in 4HBD is without precedent in anaerobic radical chemistry. It differs largely from that in enzymes, which use coenzyme B12, S-adenosylmethionine, ATP-driven electron transfer, or flavin-based electron bifurcation for this purpose.  相似文献   

12.
Ishikita H 《Biochemistry》2008,47(15):4394-4402
The redox potential of the flavin mononucleotide (FMN) hydroquinones for one-electron reduction in the Desulfovibrio vulgaris ( D. vulgaris) flavodoxin ( E sq/hq for FMNH (*)/FMNH (-)) was calculated using the crystal structure of the relevant hydroquinone form and compared to the results of the Clostridium beijerinckii ( C. beijerinckii) flavodoxin. In D. vulgaris and C. beijerinckii flavodoxins, the protein side chain causes significant downshifts of 170 and 240 mV in E sq/hq, respectively. In the C. beijerinckii flavodoxin, the E sq/hq downshift because of the protein side chain is essentially compensated by the counter influence of the protein backbone ( E sq/hq upshift of 260 mV). However, in the D. vulgaris flavodoxin, the corresponding protein backbone influence on E sq/hq is significantly small, i.e., less than half of that in the C. beijerinckii flavodoxin. In particular, there is a significant difference in the influence of the protein backbone of the so-called 60s loop region between the two flavodoxins. The E sq/hq difference can be best explained by the lower compensation of the side chain influence by the backbone influence in the D. vulgaris flavodoxin than in the C. beijerinckii flavodoxin.  相似文献   

13.
Blazyk JL  Lippard SJ 《Biochemistry》2002,41(52):15780-15794
Soluble methane monooxygenase (sMMO) from Methylococcus capsulatus (Bath) catalyzes the selective oxidation of methane to methanol, the first step in the primary catabolic pathway of methanotrophic bacteria. A reductase (MMOR) mediates electron transfer from NADH through its FAD and [2Fe-2S] cofactors to the dinuclear non-heme iron sites housed in a hydroxylase (MMOH). The structurally distinct [2Fe-2S], FAD, and NADH binding domains of MMOR facilitated division of the protein into its functional ferredoxin (MMOR-Fd) and FAD/NADH (MMOR-FAD) component domains. The 10.9 kDa MMOR-Fd (MMOR residues 1-98) and 27.6 kDa MMOR-FAD (MMOR residues 99-348) were expressed and purified from recombinant Escherichia coli systems. The Fd and FAD domains have absorbance spectral features identical to those of the [2Fe-2S] and flavin components, respectively, of MMOR. Redox potentials, determined by reductive titrations that included indicator dyes, for the [2Fe-2S] and FAD cofactors in the domains are as follows: -205.2 +/- 1.3 mV for [2Fe-2S](ox/red), -172.4 +/- 2.0 mV for FAD(ox/sq), and -266.4 +/- 3.5 mV for FAD(sq/hq). Kinetic and spectral properties of intermediates observed in the reaction of oxidized MMOR-FAD (FAD(ox)) with NADH at 4 degrees C were established with stopped-flow UV-visible spectroscopy. Analysis of the influence of pH on MMOR-FAD optical spectra, redox potentials, and NADH reaction kinetics afforded pK(a) values for the semiquinone (FAD(sq)) and hydroquinone (FAD(hq)) MMOR-FAD species and two protonatable groups near the flavin cofactor. Electron transfer from MMOR-FAD(hq) to oxidized MMOR-Fd is extremely slow (k = 1500 M(-1) s(-1) at 25 degrees C, compared to 90 s(-1) at 4 degrees C for internal electron transfer between cofactors in MMOR), indicating that cofactor proximity is essential for efficient interdomain electron transfer.  相似文献   

14.
Ravasio S  Curti B  Vanoni MA 《Biochemistry》2001,40(18):5533-5541
Glutamate synthase is a complex iron-sulfur flavoprotein that catalyzes the reductive transfer of the L-glutamine amide group to C(2) of 2-oxoglutarate, forming two molecules of L-glutamate. The bacterial enzyme is an alphabeta protomer, which contains one FAD (on the beta subunit, approximately 50 kDa), one FMN (on the alpha subunit, approximately 150 kDa), and three different Fe-S clusters (one 3Fe-4S center on the alpha subunit and two 4Fe-4S clusters at an unknown location). To address the problem of the intramolecular electron pathway, we have measured the midpoint potential values of the flavin cofactors and of the 3Fe-4S cluster of glutamate synthase in the isolated alpha and beta subunits and in the alphabeta holoenzyme. No detectable amounts of flavin semiquinones were observed during reductive titrations of the enzyme, indicating that the midpoint potential value of each flavin(ox)/flavin(sq) couple is, in all cases, significantly more negative than that of the corresponding flavin(sq)/flavin(hq) couple. Association of the two subunits to form the alphabeta protomer does not alter significantly the midpoint potential value of the FMN cofactor and of the 3Fe-4S cluster (approximately -240 and -270 mV, respectively), but it makes that of FAD some 40 mV less negative (approximately -340 mV for the beta subunit and -300 mV for FAD bound to the holoenzyme). Binding of the nonreducible NADP(+) analogue, 3-aminopyridine adenine dinucleotide phosphate, made the measured midpoint potential value of the FAD cofactor approximately 30-40 mV less negative in the isolated beta subunit, but had no effect on the redox properties of the alphabeta holoenzyme. This result correlates with the formation of a stable charge-transfer complex between the reduced flavin and the oxidized pyridine nucleotide in the isolated beta subunit, but not in the alphabeta holoenzyme. Binding of L-methionine sulfone, a glutamine analogue, had no significant effect on the redox properties of the enzyme cofactors. On the contrary, 2-oxoglutarate made the measured midpoint potential value of the 3Fe-4S cluster approximately 20 mV more negative in the isolated alpha subunit, but up to 100 mV less negative in the alphabeta holoenzyme as compared to the values of the corresponding free enzyme forms. These findings are consistent with electron transfer from the entry site (FAD) to the exit site (FMN) through the 3Fe-4S center of the enzyme and the involvement of at least one of the two low-potential 4Fe-4S centers, which are present in the glutamate synthase holoenzyme, but not in the isolated subunits. Furthermore, the data demonstrate a specific role of 2-oxoglutarate in promoting electron transfer from FAD to the 3Fe-4S cluster of the glutamate synthase holoenzyme. The modulatory role of 2-oxoglutarate is indeed consistent with the recently determined three-dimensional structure of the glutamate synthase alpha subunit, in which several polypeptide stretches are suitably positioned to mediate communication between substrate binding sites and the enzyme redox centers (FMN and the 3Fe-4S cluster) to tightly control and coordinate the individual reaction steps [Binda, C., et al. (2000) Structure 8, 1299-1308].  相似文献   

15.
Electron-transfer flavoprotein-ubiquinone oxidoreductase (ETF-QO) is an iron-sulfur flavoprotein that accepts electrons from electron-transfer flavoprotein (ETF) and reduces ubiquinone from the Q-pool. ETF-QO contains a single [4Fe-4S]2+,1+ cluster and one equivalent of FAD, which are diamagnetic in the isolated oxidized enzyme and can be reduced to paramagnetic forms by enzymatic donors or dithionite. Mutations were introduced by site-directed mutagenesis of amino acids in the vicinity of the iron-sulfur cluster of Rhodobacter sphaeroides ETF-QO. Y501 and T525 are equivalent to Y533 and T558 in the porcine ETF-QO. In the porcine protein, these residues are within hydrogen-bonding distance of the Sgamma of the cysteine ligands to the iron-sulfur cluster. Y501F, T525A, and Y501F/T525A substitutions were made to determine the effects on midpoint potential, activity, and EPR spectral properties of the cluster. The integrity of the mutated proteins was confirmed by optical spectra, EPR g-values, and spin-lattice relaxation rates, and the cluster to flavin point-dipole distance was determined by relaxation enhancement. Potentiometric titrations were monitored by changes in the CW EPR signals of the cluster and semiquinone. Single mutations decreased the midpoint potentials of the iron-sulfur cluster from +37 mV for wild type to -60 mV for Y501F and T525A and to -128 mV for Y501F/T525A. Lowering the midpoint potential resulted in a decrease in steady-state ubiquinone reductase activity and in ETF semiquinone disproportionation. The decrease in activity demonstrates that reduction of the iron-sulfur cluster is required for activity. There was no detectable effect of the mutations on the flavin midpoint potentials.  相似文献   

16.
The oxidation-reduction potential of p-hydroxybenzoate hydroxylase (4-hydroxybenzoate, NADPH: oxygen oxidoreductase (3-hydroxylating), EC 1.14.13.2) from Pseudomonas fluorescens has been measured in the presence and absence of p-hydroxybenzoate using spectrocoulometry. The native enzyme demonstrated a two-electron midpoint potential of -129 mV during the initial reductive titration. The midpoint potential observed during subsequent oxidative and reductive titrations was -152 mV. This marked hysteresis is proposed to arise from the oxidation and reduction of the known air-sensitive thiol group on the enzyme (Van Berkel, W.J.H. and Müller, F. (1987) Eur. J. Biochem. 167, 35-46). Redox titrations of the enzyme in the presence of substrate showed a two-electron midpoint potential of -177 mV. No spectral or electrochemical evidence for the thermodynamic stabilization of any flavin semiquinone was observed in the titrations performed. These data show that the affinity of the apoenzyme for the hydroquinone form of FAD is 150-fold greater than for the oxidized flavin and that the substrate is bound to the reduced enzyme with a 3-fold lower affinity than to the oxidized enzyme. These data are consistent with the view that the stimulatory effect of substrate binding on the rate of enzyme reduction by NADPH is due to the respective geometries of the bound FAD and NADPH rather than to a large perturbation of the oxidation-reduction potential of the bound flavin coenzyme.  相似文献   

17.
Enzymology of butyrate formation by Butyrivibrio fibrisolvens.   总被引:2,自引:0,他引:2       下载免费PDF全文
Butyrivibrio fibrisolvens is a major butyrate-forming species in the bovine and ovine rumen. The enzymology of butyrate formation from pyruvate was investigated in cell-free extracts of B. fibrisolvens D1. Pyruvate owas oxidized to acetylcoenzyme A (CoA) in the presence of CoA.SH and benzyl viologen or flavin nucleotides. The bacterium uses thiolase, beta-hydroxybutyryl-CoA dehydrogenase, crotonase, and crotonyl-CoA reductase to form butyryl-CoA from acetyl-CoA. Reduction of acetoacetyl-CoA to beta-hydroxybutyryl-CoA was faster with NADH than with NADPH. Crotonyl-CoA was reduced to butyryl-CoA by NADH, but not by NADPH, only in the presence of flavin nucleotides. Reduction of flavin nucleotides by NADH was much slower than the flavin-dependent reduction of crotonyl-CoA. This indicates that flavoproteins rather than free flavin participated in the reduction of crotonyl-CoA. Butyryl-CoA was converted to butyrate by phosphate butyryl transferase and butyrate kinase.  相似文献   

18.
The side chain of aspartate 95 in flavodoxin from Desulfovibrio vulgaris provides the closest negative charge to N(1) of the bound FMN in the protein. Site-directed mutagenesis was used to substitute alanine, asparagine, or glutamate for this amino acid to assess the effect of this charge on the semiquinone/hydroquinone redox potential (E(1)) of the FMN cofactor. The D95A mutation shifts the E(1) redox potential positively by 16 mV, while a negative shift of 23 mV occurs in the oxidized/semiquinone midpoint redox potential (E(2)). The crystal structures of the oxidized and semiquinone forms of this mutant are similar to the corresponding states of the wild-type protein. In contrast to the wild-type protein, a further change in structure occurs in the D95A mutant in the hydroquinone form. The side chain of Y98 flips into an energetically more favorable edge-to-face interaction with the bound FMN. Analysis of the structural changes in the D95A mutant, taking into account electrostatic interactions at the FMN binding site, suggests that the pi-pi electrostatic repulsions have only a minor contribution to the very low E(1) redox potential of the FMN cofactor when bound to apoflavodoxin. Substitution of D95 with glutamate causes only a slight perturbation of the two one-electron redox potentials of the FMN cofactor. The structure of the D95E mutant reveals a large movement of the 60-loop (residues 60-64) away from the flavin in the oxidized structure. Reduction of this mutant to the hydroquinone causes the conformation of the 60-loop to revert back to that occurring in the structures of the wild-type protein. The crystal structures of the D95E mutant imply that electrostatic repulsion between a carboxylate on the side chain at position 95 and the phenol ring of Y98 prevents rotation of the Y98 side chain to a more energetically favorable conformation as occurs in the D95A mutant. Replacement of D95 with asparagine has no effect on E(2) but causes E(1) to change by 45 mV. The D95N mutant failed to crystallize. The K(d) values of the protein FMN complex in all three oxidation-reduction states differ from those of the wild-type complexes. Molecular modeling showed that the conformational energy of the protein changes with the redox state, in qualitative agreement with the observed changes in K(d), and allowed the electrostatic interactions between the FMN and the surrounding groups on the protein to be quantified.  相似文献   

19.
The two flavodoxins (YkuN and YkuP) from Bacillus subtilis have been cloned, overexpressed in Escherichia coli and purified. DNA sequencing, mass spectrometry, and flavin-binding properties showed that both YkuN and YkuP were typical short-chain flavodoxins (158 and 151 amino acids, respectively) and that an error in the published B. subtilis genome sequence had resulted in an altered reading frame and misassignment of YkuP as a long-chain flavodoxin. YkuN and YkuP were expressed in their blue (neutral semiquinone) forms and reoxidized to the quinone form during purification. Potentiometry confirmed the strong stabilization of the semiquinone form by both YkuN and YkuP (midpoint reduction potential for oxidized/semiquinone couple = -105 mV/-105 mV) with respect to the hydroquinone (midpoint reduction potential for semiquinone/hydroquinone couple = -382 mV/-377 mV). Apoflavodoxin forms were generated by trichloroacetic acid treatment. Circular dichroism studies indicated that flavin mononucleotide (FMN) binding led to considerable structural rearrangement for YkuP but not for YkuN. Both apoflavodoxins bound FMN but not riboflavin avidly, as expected for short-chain flavodoxins. Structural stability studies with the chaotrope guanidinium chloride revealed that there is moderate destabilization of secondary and tertiary structure on FMN removal from YkuN, but that YkuP apoflavodoxin has similar (or slightly higher) stability compared to the holoprotein. Differential scanning calorimetry reveals further differences in structural stability. YkuP has a lower melting temperature than YkuN, and its endotherm is composed of a single transition, while that for YkuN is biphasic. Optical and fluorimetric titrations with oxidized flavodoxins revealed strong affinity (K(d) values consistently <5 microM) for their potential redox partner P450 BioI, YkuN showing tighter binding. Stopped-flow reduction studies indicated that the maximal electron-transfer rate (k(red)) to fatty acid-bound P450 BioI occurs from YkuN and YkuP at approximately 2.5 s(-1), considerably faster than from E. coli flavodoxin. Steady-state turnover with YkuN or YkuP, fatty acid-bound P450 BioI, and E. coli NADPH-flavodoxin reductase indicated that both flavodoxins supported lipid hydroxylation by P450 BioI with turnover rates of up to approximately 100 min(-1) with lauric acid as substrate. Interprotein electron transfer is a likely rate-limiting step. YkuN and YkuP supported monohydroxylation of lauric acid and myristic acid, but secondary oxygenation of the primary product was observed with both palmitic acid and palmitoleic acid as substrates.  相似文献   

20.
Cell extracts of butyrate-forming clostridia have been shown to catalyze acetyl-coenzyme A (acetyl-CoA)- and ferredoxin-dependent formation of H2 from NADH. It has been proposed that these bacteria contain an NADH:ferredoxin oxidoreductase which is allosterically regulated by acetyl-CoA. We report here that ferredoxin reduction with NADH in cell extracts from Clostridium kluyveri is catalyzed by the butyryl-CoA dehydrogenase/Etf complex and that the acetyl-CoA dependence previously observed is due to the fact that the cell extracts catalyze the reduction of acetyl-CoA with NADH via crotonyl-CoA to butyryl-CoA. The cytoplasmic butyryl-CoA dehydrogenase complex was purified and is shown to couple the endergonic reduction of ferredoxin (E0′ = −410 mV) with NADH (E0′ = −320 mV) to the exergonic reduction of crotonyl-CoA to butyryl-CoA (E0′ = −10 mV) with NADH. The stoichiometry of the fully coupled reaction is extrapolated to be as follows: 2 NADH + 1 oxidized ferredoxin + 1 crotonyl-CoA = 2 NAD+ + 1 ferredoxin reduced by two electrons + 1 butyryl-CoA. The implications of this finding for the energy metabolism of butyrate-forming anaerobes are discussed in the accompanying paper.  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