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 920 毫秒
1.
Flavodoxin from Desulfovibrio vulgaris (Hildenborough) has been expressed at a high level (3-4% soluble protein) in Escherichia coli by subcloning a minimal insert carrying the gene behind the tac promoter of plasmid pDK6. The recombinant protein was readily isolated and its properties were shown to be identical to those of the wild-type protein obtained directly from D. vulgaris, with the exception that the recombinant protein lacks the N-terminal methionine residue. Detailed measurements of the redox potentials of this flavodoxin are reported for the first time. The redox potential, E2, for the couple oxidized flavodoxin/flavodoxin semiquinone at pH 7.0 is -143 mV (25 degrees C), while the value for the flavodoxin semiquinone/flavodoxin hydroquinone couple (E1) at the same pH is -440 mV. The effects of pH on the observed potentials were examined; E2 varies linearly with pH (slope = -59 mV), while E1 is independent of pH at high pH values, but below pH 7.5 the potential becomes less negative with decreasing pH, indicating a redox-linked protonation of the flavodoxin hydroquinone. D. vulgaris apoflavodoxin binds FMN very tightly, with a value of 0.24 nM for the dissociation constant (Kd) at pH 7.0 and 25 degrees C, similar to that observed with other flavodoxins. In addition, the apoflavodoxin readily binds riboflavin (Kd = 0.72 microM; 50 mM sodium phosphate, pH 7.0, 5 mM EDTA at 25 degrees C) and the complex is spectroscopically very similar to that formed with FMN. The redox potentials for the riboflavin complex were determined at pH 6.5 (E1 = -262 mV, E2 = -193 mV; 25 degrees C) and are discussed in the light of earlier proposals that charge/charge interactions between different parts of the flavin hydroquinone play a crucial role in determining E1 in flavodoxin.  相似文献   

2.
The electrostatically stabilized complex between Anabaena variabilis ferredoxin--NADP+ reductase and Azotobacter vinelandii flavodoxin has been covalently cross-linked by treatment with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide. The covalent complex exhibits a molecular mass and FMN/FAD content consistent with that expected for a 1:1 stoichiometry of the two flavoproteins. Immunochemical cross-reactivity is exhibited by the covalent complex with rabbit antisera prepared separately against each protein. The complex retains NADPH-ferricyanide diaphorase activity although the Km for ferricyanide is increased twofold and the turnover number is decreased by a factor of two when compared to native reductase. NADPH-cytochrome-c reductase activity of the complex is observed at a level that is quite similar to that determined at saturating concentrations of flavodoxin, while it is only 1-2% of that exhibited by the reductase in the presence of ferredoxin. No stimulation of cytochrome-c reductase activity is observed on adding ferredoxin to the cross-linked complex. Stopped-flow data show that covalent cross-linking of the flavodoxin to the reductase reduces the rate of electron transfer from its semiquinone form to cytochrome c by a factor of 60. Anaerobic titrations of the reduced complex with NADP+ show the semiquinone/quinol couple of the flavodoxin is increased 100 mV relative to the free form and the quinone/quinol couple of complexed ferredoxin-NADP+ reductase is increased by only 25 mV, relative to the free protein. Addition of NADPH to the cross-linked complex reduces the FAD of the reductase as well as the FMN moiety of flavodoxin to a mixture of semiquinone and quinol forms.  相似文献   

3.
The oxidation-reduction potentials of ferredoxin-NADP+ reductase and flavodoxin from the cyanobacterium Anabaena PCC 7119 were determined by potentiometry. The potentials at pH 7 for the oxidized flavodoxin/flavodoxin semiquinone couple (E2) and the flavodoxin semiquinone/hydroquinone couple (E1) were -212 mV and -436 mV, respectively. E1 was independent of pH above about pH 7, but changed by approximately -60 mV/pH below about pH 6, suggesting that the fully reduced protein has a redox-linked pKa at about 6.1, similar to those of certain other flavodoxins. E2 varied by -50 mV/pH in the range pH 5-8. The redox potential for the two-electron reduction of ferredoxin-NADP+ reductase was -344 mV at pH 7 (delta Em = -30 mV/pH). In the 1:1 electrostatic complex of the two proteins titrated at pH 7, E2 was shifted by +8 mV and E1 was shifted by -25 mV; the shift in potential for the reductase was +4 mV. The potentials again shifted following treatment of the electrostatic complex with a carbodiimide, to covalently link the two proteins. By comparison with the separate proteins at pH 7, E2 for flavodoxin shifted by -21 mV and E1 shifted by +20 mV; the reductase potential shifted by +2 mV. The potentials of the proteins in the electrostatic and covalent complexes showed similar pH dependencies to those of the individual proteins. Qualitatively similar changes occurred when ferredoxin-NADP+ reductase from Anabaena variabilis was complexed with flavodoxin from Azotobacter vinelandii. The shifts in redox potential for the complexes were used with previously determined values for the dissociation constant (Kd) of the electrostatic complex of the two oxidised proteins, in order to estimate Kd values for the interaction of the different redox forms of the proteins. The calculations showed that the electrostatic complexes, formed when the proteins differ in their redox states, are stronger than those formed when both proteins are fully oxidized or fully reduced.  相似文献   

4.
The gene for the electron-transfer protein flavodoxin has been cloned from Megasphaera elsdenii using the polymerase chain reaction. The recombinant gene was sequenced, expressed in an Escherichia coli expression system, and the recombinant protein purified and characterized. With the exception of an additional methionine residue at the N-terminus, the physico-chemical properties of the protein, including its optical spectrum and oxidation-reduction properties, are very similar to those of native flavodoxin. A site-directed mutant, E60Q, was made to investigate the effects of removing the negatively charged group that is nearest to N(1) of the bound FMN. The absorbance maximum in the visible region of the bound flavin moves from 446 to 453 nm. The midpoint oxidation-reduction potential at pH 7 for reduction of oxidized flavodoxin to the semiquinone E2 becomes more negative, decreasing from -114 to -242 mV; E1, the potential for reduction of semiquinone to the hydroquinone, becomes less negative, increasing from -373 mV to -271 mV. A redox-linked pKa associated with the hydroquinone is decreased from 5.8 to < or = 4.3. The spectra of the hydroquinones of wild-type and mutant proteins depend on pH (apparent pKa values of 5.8 and < or = 5.2, respectively). The complexes of apoprotein and all three redox forms of FMN are much weaker for the mutant, with the greatest effect occurring when the flavin is in the semiquinone form. These results suggest that glutamate 60 plays a major role in control of the redox properties of M. elsdenii flavodoxin, and they provide experimental support to an earlier proposal that the carboxylate on its side-chain is associated with the redox-linked pKa of 5.8 in the hydroquinone.  相似文献   

5.
The flavodoxins of the Azotobacter vinelandii wild-type and a mutant strain TZN 200 have been studied. Although the primary structure of the two proteins is the same, the ability of the mutant flavodoxin to donate electrons to nitrogenase is reduced by 75%. One reason may be the raised mid-point potential of -435 mV for the semiquinone/hydroquinone couple in the mutant flavodoxin. The respective redox potential for the wild-type flavodoxin was found to be -480 mV. As shown by paper chromatography and light absorption spectroscopy, the structure of FMN is modified in the TZN 200 flavodoxin.  相似文献   

6.
Structural and chemical properties of a flavodoxin from Anabaena PCC 7119   总被引:1,自引:0,他引:1  
Structural and chemical properties of a flavodoxin from Anabaena PCC 7119 are described. The first 36 residues of the amino-terminal amino acid sequence have been determined and show extensive homology with flavodoxins isolated from other sources. Anabaena flavodoxin exhibits a net negative change (-3) in the helix-1 segment as found with other cyanobacterial flavodoxins Synechococcus 6301 (Anacystis nidulans) and Nostoc MAC, but in contrast to the net positive charge found in this region in the case of flavodoxins isolated from nitrogen-fixing bacteria (Azotobacter and Klebsiella). The FMN cofactor can be reversibly resolved from the apoprotein by trichloroacetic acid treatment. Apoflavodoxin, thus prepared, binds FMN with a Kd value of 0.1 nM and binds riboflavin with a decreased affinity (Kd = 5 microM) at pH 7.2. The apoprotein is stable in dilute solutions at pH values around 7 but readily denatures at pH 8 as judged from loss in flavin-binding ability and by ultraviolet circular dichroism spectroscopy. Oxidation-reduction potential studies at pH values of 7 and 8 show OX/SQ couples of -195 mV and -255 mV, respectively, and show SQ/HQ couples of -390 mV and -418 mV, respectively. From these data, the binding constant for the FMN semiquinone is calculated to be approx. 5-fold tighter and the binding of the FMN hydroquinone is approx. 10(5)-fold weaker than that of the oxidized FMN to the apoprotein. Anabaena flavodoxin functions as an effective mediator of electron transfer from ferredoxin-NADP(+)-reductase to cytochrome c with a turnover number [4.5-5) x 10(3) min-1); a values similar to that determined for Anabaena ferredoxin. The flavodoxin binds tightly to the reductase with Kd values of 6.4 and 8.5 microM at pH values of 7.0 and 8.0, respectively.  相似文献   

7.
Azotobacter vinelandii flavodoxin II serves as a physiological reductant of nitrogenase, the enzyme system mediating biological nitrogen fixation. Wildtype A. vinelandii flavodoxin II was electrochemically and crystallographically characterized to better understand the molecular basis for this functional role. The redox properties were monitored on surfactant‐modified basal plane graphite electrodes, with two distinct redox couples measured by cyclic voltammetry corresponding to reduction potentials of ?483 ± 1 mV and ?187 ± 9 mV (vs. NHE) in 50 mM potassium phosphate, 150 mM NaCl, pH 7.5. These redox potentials were assigned as the semiquinone/hydroquinone couple and the quinone/semiquinone couple, respectively. This study constitutes one of the first applications of surfactant‐modified basal plane graphite electrodes to characterize the redox properties of a flavodoxin, thus providing a novel electrochemical method to study this class of protein. The X‐ray crystal structure of the flavodoxin purified from A. vinelandii was solved at 1.17 Å resolution. With this structure, the native nitrogenase electron transfer proteins have all been structurally characterized. Docking studies indicate that a common binding site surrounding the Fe‐protein [4Fe:4S] cluster mediates complex formation with the redox partners Mo‐Fe protein, ferredoxin I, and flavodoxin II. This model supports a mechanistic hypothesis that electron transfer reactions between the Fe‐protein and its redox partners are mutually exclusive.  相似文献   

8.
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   

9.
Reductive titration curves of flavodoxin from Desulfovibrio vulgaris displayed two one-electron steps. The redox potential E-2 for the couple oxidized flavodoxin/flavodoxin semiquinone was determined by direct titration with dithionite. E-2 was -149 plus or minus 3 mV (pH 7.78, 25 degrees C). The redox potential E-1 for the couple flavodoxin semiquinone/fully reduced flavodoxin was deduced from the equilibrium concentration of these species in the presence of hydrogenase and H-2. E-1 was -438 plus or minus 8 mV (pH 7.78, 25 degrees C). Light-absorption and fluorescence spectra of flavodoxin in its three redox states have been recorded. Both the rate and extent of reduction of flavodoxin semiguinone with dithionite were found to depend on pH. An equilibrium between the semiquinone and hydroquinone forms occurred at pH values close to the neutrality, even in the presence of a large excess of dithionite, suggesting an ionization in fully reduced flavodoxin with a pK-a = 6.6. The association constants K for the three FMN redox forms with the apoprotein were deduced from the value of K (K = 8 times 10-7 M-1) measured with oxidized EMN at pH 7.0. Oxidized flavodoxin was found to comproportionate with the fully reduced protein (k-comp = 4.3 times 10-3 M-1 times s-1, pH 9.0, 22 degrees C) and with reduced free FMN (K-comp = 44 M-1 times s-1, pH 8.1, 20 degrees C). Fast oxidation of reduced flavodoxin occurred in the presence of O-2. Slower oxidation of semiquinone was dependent on pH in a drastic way.  相似文献   

10.
Characterization of three different flavodoxins from Azotobacter vinelandii   总被引:7,自引:0,他引:7  
The flavodoxins from Azotobacter vinelandii cells grown N2-fixing and from cells grown on NH4OAc have been purified and characterized. The purified flavodoxins from these cells are a mixture of three different flavodoxins (Fld I, II, III) with different primary structures. The three proteins were separated by fast protein liquid chromatography; Fld I eluted at 0.38 M KCl, Fld II at 0.43 M KCl and Fld III at 0.45 M KCl. The most striking difference between the three flavodoxins was the midpoint potential (pH 7.0, 25 degrees C) of the semiquinone/hydroquinone couple, which was -320 mV for Fld I and -500 mV for the other two flavodoxins (Fld II and Fld III). All three flavodoxins were present in cells grown on NH4OAc. In cells grown on N2 as N source only Fld I and Fld II were found. The concentration of Fld II was 10-fold higher in N2-fixing cells than in cells grown on NH4OAc. Evidence has been obtained that Fld II is involved in electron transport to nitrogenase. As will be discussed, our observation that preparations of Azotobacter flavodoxin are heterogeneous, has consequences for the published data.  相似文献   

11.
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   

12.
Pulse-radiolysis experiments were performed on solutions containing methyl or benzyl viologen and flavodoxin. Viologen radicals are formed after the pulse. The kinetics of the reaction of these radicals with flavodoxin were studied. The kinetics observed depend strongly on the concentration of oxidized viologen. Therefore one must conclude that a relatively stable intermediate is formed after the reduction of flavodoxin. The midpoint potential of the intermediate state is -(480 +/- 30) mV, and is hardly dependent on the pH between 7 and 9.2. Due to a conformational change (k2 approximately equal to 10(5)S-1) the intermediate state decays to the stable semiquinone form of flavodoxin. The delta G of the conformational change at pH 8 is about 29 kJ mol -1 (0.3 eV). This means that the upper limit for the pK of N-5 in the semiquinone form will be 13. The activation energy of the conformational change is 43 kJ mol -1 (0.45 eV). The reaction between methyl viologen radicals and the semiquinone of flavodoxin can be described by a normal bimolecular reaction. The reaction is diffusion-controlled with a forward rate constant of (7 +/- 1) X 10(8) M -1S -1 (pH 8, I = 55 mM). The midpoint potential of the semiquinone/hydroquinone was found to be -(408 +/- 5) mV. A consequence of the intermediate state is that flavodoxin (Fld) could be reduced by a two-electron process, the midpoint potential of which should be located between -440 mV less than Em (Fld/FldH-) less than -290 mV. The exact value will depend on the delta G of the conformational change between the fully reduced flavodoxin with its structure in the oxidized form and the fully reduced flavodoxin with its structure in the hydroquinone form. The conditions are discussed under which flavodoxin could behave as a two-electron donor.  相似文献   

13.
Flavodoxin in the hydroquinone state acts as an electron donor to nitrogenase in several nitrogen-fixing organisms. The mid-point potentials for the oxidized-semiquinone and semiquinone-hydroquinone couples of flavodoxins isolated from facultative anaerobe Klebsiella pneumoniae (nifF-gene product, KpFld) and the obligate aerobe Azotobacter chroococcum (AcFld) were determined as a function of pH. The mid-point potentials of the semiquinone-hydroquinone couples of KpFld and AcFld are essentially independent of pH over the range pH 7-9, being -422 mV and -522 mV (normal hydrogen electrode) at pH 7.5 respectively. The mid-point potentials of the quinone-semiquinone couples at pH 7.5 are -200 mV (KpFld) and -133 mV (AcFld) with delta Em/pH of -65 +/- 4 mV (KpFld) and -55 +/- 2 mV (AcFld) over the range pH 7.0-9.5. This indicates that reduction of the quinone is coupled to protonation to yield a neutral semiquinone. The significance of these values with respect to electron transport to nitrogenase is discussed. The amino acid compositions, the N- and C-terminal amino acid sequences and the u.v.-visible spectra of KpFld and AcFld were determined and are compared with published data for flavodoxins isolated from Azotobacter vinelandii.  相似文献   

14.
The gene coding for the flavodoxin protein from Desulfovibrio vulgaris (Hildenborough) has been identified, cloned, and sequenced. DNA fragments containing the flavodoxin gene were identified by hybridization of a mixed synthetic heptadecanucleotide probe to Southern blots of SalI-digested genomic DNA. The nucleotide sequences of the probe were derived from the published protein primary structure (Dubourdieu, M., LeGall, J., and Fox, J. L. (1973) Biochem. Biophys. Res. Commun. 52, 1418-1425). The same oligonucleotide probe was used to screen libraries (in pUC19) containing size-selected SalI fragments. One recombinant, carrying a 1.6-kilobase (kb) insert which strongly hybridizes to the probe, was found to contain a nucleotide sequence which codes for the first 104 residues of the amino-terminal portion of the flavodoxin protein sequence but lacked the remainder of the gene. Therefore, a PstI restriction fragment from this clone was used as a probe to isolate the entire gene from a partial Sau3AI library in Charon 35. Of the plaques which continued to hybridize strongly to this probe through repeated screenings, one recombinant, containing a 16-kb insert, was further characterized. The entire flavodoxin gene was localized within a 1.4-kb XhoI-SacI fragment of this clone. The complete nucleotide sequence of the structural gene for the flavodoxin protein from Desulfovibrio vulgaris and flanking sequences which may include promoter and regulatory sequences are reported here. The cloned flavodoxin gene was placed behind the hybrid tac promoter for overexpression of the protein in Escherichia coli. Modification to the 5'-end of the gene, including substitutions at the second codon, were required to obtain high levels of expression. The expressed recombinant flavodoxin protein is isolated from E. coli cells as the holoprotein with physical and spectral properties similar to the protein isolated from D. vulgaris. To our knowledge, this is the first example of the expression of a foreign flavodoxin gene in E. coli using recombinant DNA methods.  相似文献   

15.
Mammalian nitric-oxide synthases are large modular enzymes that evolved from independently expressed ancestors. Calmodulin-controlled isoforms are signal generators; calmodulin activates electron transfer from NADPH through three reductase domains to an oxygenase domain. Structures of the reductase unit and its homologs show FMN and FAD in contact but too isolated from the protein surface to permit exit of reducing equivalents. To study states in which FMN/heme electron transfer is feasible, we designed and produced constructs including only oxygenase and FMN binding domains, eliminating strong internal reductase complex interactions. Constructs for all mammalian isoforms were expressed and purified as dimers. All synthesize NO with peroxide as the electron donor at rates comparable with corresponding oxygenase constructs. All bind cofactors nearly stoichiometrically and have native catalytic sites by spectroscopic criteria. Modest differences in electrochemistry versus independently expressed heme and FMN binding domains suggest interdomain interactions. These interactions can be convincingly demonstrated via calmodulin-induced shifts in high spin ferriheme EPR spectra and through mutual broadening of heme and FMNH. radical signals in inducible nitric-oxide synthase constructs. Blue neutral FMN semiquinone can be readily observed; potentials of one electron couple (in inducible nitric-oxide synthase oxygenase FMN, FMN oxidized/semiquinone couple = +70 mV, FMN semiquinone/hydroquinone couple = -180 mV, and heme = -180 mV) indicate that FMN is capable of serving as a one electron heme reductant. The construct will serve as the basis for future studies of the output state for NADPH derived reducing equivalents.  相似文献   

16.
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   

17.
Midpoint reduction potentials for the flavin cofactors in human NADPH-cytochrome P450 oxidoreductase were determined by anaerobic redox titration of the diflavin (FAD and FMN) enzyme and by separate titrations of its isolated FAD/NADPH and FMN domains. Flavin reduction potentials are similar in the isolated domains (FAD domain E(1) [oxidized/semiquinone] = -286 +/- 6 mV, E(2) [semiquinone/reduced] = -371 +/- 7 mV; FMN domain E(1) = -43 +/- 7 mV, E(2) = -280 +/- 8 mV) and the soluble diflavin reductase (E(1) [FMN] = -66 +/- 8 mV, E(2) [FMN] = -269 +/- 10 mV; E(1) [FAD] = -283 +/- 5 mV, E(2) [FAD] = -382 +/- 8 mV). The lack of perturbation of the individual flavin potentials in the FAD and FMN domains indicates that the flavins are located in discrete environments and that these environments are not significantly disrupted by genetic dissection of the domains. Each flavin titrates through a blue semiquinone state, with the FMN semiquinone being most intense due to larger separation (approximately 200 mV) of its two couples. Both the FMN domain and the soluble reductase are purified in partially reduced, colored form from the Escherichia coli expression system, either as a green reductase or a gray-blue FMN domain. In both cases, large amounts of the higher potential FMN are in the semiquinone form. The redox properties of human cytochrome P450 reductase (CPR) are similar to those reported for rabbit CPR and the reductase domain of neuronal nitric oxide synthase. However, they differ markedly from those of yeast and bacterial CPRs, pointing to an important evolutionary difference in electronic regulation of these enzymes.  相似文献   

18.
19.
Using pBR322- and pUC-derived plasmid vectors, a homologous (Escherichia coli native esterase) and three heterologous proteins (human interleukin-2, human interleukin-6, and Zymomonas levansucrase) were synthesized in E. coli IC2015(recA::lacZ) and GY4786 (sfiA::lacZ) strains. Via time-course measurement of beta-galactosidase activity in each recombinant culture, the SOS induction was estimated in detail and the results were systematically compared. In recombinant E. coli, the SOS response did not happen either with the recombinant insert-negative plasmid backbone alone or the expression vectors containing the homologous gene. Irrespective of gene expression level and toxic activity of synthesized foreign proteins, the SOS response was induced only when the heterologous genes were expressed using a particular plasmid vector, indicating strong dependence on the recombinant gene clone and the selection of a plasmid vector system. It is suggested that in recombinant E. coli the SOS response (i.e., activation of recA expression and initial sfiA expression) may be related neither to metabolic burden nor toxic cellular event(s) by synthesized heterologous protein, but may be provoked by foreign gene-specific interaction between a foreign gene and a plasmid vector. Unlike in E. coli XL1-blue(recA(-)) strains used, all expression vectors encoding each of the three heterologous proteins were multimerized in E. coli IC2015 strains in the course of cultivation, whereas the expression vectors containing the homologous gene never formed the plasmid multimers. The extent of multimerization was also dependent on a foreign gene insert in the expression vector. As a dominant effect of the SOS induction, recombinant plasmid vectors used for heterologous protein expression appear to significantly form various multimers in the recA(+) E. coli host.  相似文献   

20.
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   

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