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.
Luminous bacteria contain several species of flavin reductases, which catalyze the reduction of FMN using NADH and/or NADPH as a reductant. The reduced FMN (i.e. FMNH(2)) so generated is utilized along with a long-chain aliphatic aldehyde and molecular oxygen by luciferase as substrates for the bioluminescence reaction. In this report, the general properties of luciferases and reductases from luminous bacteria are briefly summarized. Earlier and more recent studies demonstrating the direct transfer of FMNH(2) from reductases to luciferase are surveyed. Using reductases and luciferases from Vibrio harveyi and Vibrio fischeri, two mechanisms were uncovered for the direct transfer of reduced flavin cofactor and reduced flavin product of reductase to luciferase. A complex of an NADPH-specific reductase (FRP(Vh)) and luciferase from V. harveyi has been detected in vitro and in vivo. Both constituent enzymes in such a complex are catalytically active. The reduction of FRP(Vh)-bound FMN cofactor by NADPH is reversible, allowing the cellular contents of NADP(+) and NADPH as a factor for the regulation of the production of FMNH(2) by FRP(Vh) for luciferase bioluminescence. Other regulations of the activity coupling between reductase and luciferase are also discussed.  相似文献   

2.
Jeffers CE  Nichols JC  Tu SC 《Biochemistry》2003,42(2):529-534
A direct transfer of the reduced flavin mononucleotide (FMNH(2)) cofactor of Vibrio harveyi NADPH:FMN oxidoreductase (FRP) to luciferase for the coupled bioluminescence reaction has been indicated by recent kinetic studies [Lei, B., and Tu, S.-C. (1998) Biochemistry 37, 14623-14629; Jeffers, C., and Tu, S.-C. (2001) Biochemistry 40, 1749-1754]. For such a mechanism, a complex formation of luciferase with FRP is essential, but until now, no evidence for such a complex has been reported. In this work, FRP was labeled at 1:1 molar ratio with the fluorophore eosin. The labeled enzyme was about 30% active in either the reductase single-enzyme or the luciferase-coupled assay. The labeled FRP in either the holo- or apoenzyme form was similar to the native FRP in undergoing a monomer-dimer equilibrium. By measuring the steady-state fluorescence anisotropy of eosin-labeled FRP, it was shown that luciferase formed a complex at 1:1 molar ratio with the monomer of either the apoenzyme or the holoenzyme form of FRP with K(d) values of 7 and 11 microM, respectively. Neither the holo- nor the apoenzyme of the labeled FRP in the dimeric form was effective in complexing with luciferase. At maximal in vivo bioluminescence, the V. harveyi cellular contents of luciferase and FRP were estimated to be 172 and 3 microM, respectively. The vast majority of FRP would be trapped in the luciferase/FRP complex. Plausible physiological significance of such a finding is discussed.  相似文献   

3.
Wang H  Lei B  Tu SC 《Biochemistry》2000,39(26):7813-7819
Luminous bacteria contain three types of NAD(P)H-FMN oxidoreductases (flavin reductases) with different pyridine nucleotide specificities. Among them, the NADPH-specific flavin reductase from Vibrio harveyi exhibits a uniquely high preference for NADPH. In comparing the substrate specificity, crystal structure, and primary sequence of this flavin reductase with other structurally related proteins, we hypothesize that the conserved Arg203 residue of this reductase is critical to the specific recognition of NADPH. The mutation of this residue to an alanine resulted in only small changes in the binding and reduction potential of the FMN cofactor, the K(m) for the FMN substrate, and the k(cat). In contrast, the K(m) for NADPH was increased 36-fold by such a mutation. The characteristic perturbation of the FMN cofactor absorption spectrum upon NADP(+) binding by the wild-type reductase was abolished by the same mutation. While the k(cat)/K(m,NADPH) was reduced from 1990 x 10(5) to 46 x 10(5) M(-1) min(-1) by the mutation, the mutated variant showed a k(cat)/K(m,NADH) of 4 x 10(5) M(-1) min(-1), closely resembling that of the wild-type reductase. The deuterium isotope effects (D)V and (D)(V/K) for (4R)-[4-(2)H]-NADPH were 1.7 and 1.4, respectively, for the wild-type reductase but were increased to 3.8 and 4.0, respectively, for the mutated variant. Such a finding indicates that the rates of NADPH and NADP(+) dissociation in relation to the isotope-sensitive redox steps were both increased as a result of the mutation. These results all provide support to the critical role of the Arg203 in the specific recognition and binding of NADPH.  相似文献   

4.
Jawanda N  Ahmed K  Tu SC 《Biochemistry》2008,47(1):368-377
Vibrio harveyi luciferase and flavin reductase FRP are, together, a two-component monooxygenase couple. The reduced flavin mononucleotide (FMNH2) generated by FRP must be supplied, through either free diffusion or direct transfer, to luciferase as a substrate. In contrast, single-component bifunctional monooxygenases each contains a bound flavin cofactor and does not require any flavin addition to facilitate catalysis. In this study, we generated and characterized a novel fusion enzyme, FRP-alphabeta, in which FRP was fused to the luciferase alpha subunit. Both FRP and luciferase within FRP-alphabeta were catalytically active. Kinetic properties characteristic of a direct transfer of FMNH2 cofactor from FRP to luciferase in a FRP:luciferase noncovalent complex were retained by FRP-alphabeta. At submicromolar levels, FRP-alphabeta was significantly more active than an equal molar mixture of FRP and luciferase in coupled bioluminescence without FMN addition. Importantly, FRP-alphabeta gave a higher total quantum output without than with exogenously added FMN. Moreover, effects of increasing concentrations of oxygen on light intensity were investigated using sub-micromolar enzymes, and results indicated that the bioluminescence produced by FRP-alphabeta without added flavin was derived from direct transfer of reduced flavin whereas bioluminescence from a mixture of FRP and luciferase with or without exogenously added flavin relied on free-diffusing reduced flavin. Therefore, the overall catalytic reaction of FRP-alphabeta without any FMN addition closely mimics that of a single-component bifunctional monooxygenase. This fusion enzyme approach could be useful to other two-component monooxygenases in enhancing the enzyme efficiencies under conditions hindering reduced flavin delivery. Other potential utilities of this approach are discussed.  相似文献   

5.
Jeffers CE  Tu SC 《Biochemistry》2001,40(6):1749-1754
It is believed that the reduced FMN substrate required by luciferase from luminous bacteria is provided in vivo by NAD(P)H-FMN oxidoreductases (flavin reductases). Our earlier kinetic study indicates a direct flavin cofactor transfer from Vibrio harveyi NADPH-preferring flavin reductase P (FRP(H)) to the luciferase (L(H)) from the same bacterium in the in vitro coupled luminescence reaction. Kinetic studies were carried out in this work to characterize coupled luminescence reactions using FRP(H) and the Vibrio fischeri NAD(P)H-utilizing flavin reductase G (FRG(F)) in combination with L(H) or luciferase from V. fischeri (L(F)). Comparisons of K(m) values of reductases for flavin and pyridine nucleotide substrates in single-enzyme and luciferase-coupled assays indicate a direct transfer of reduced flavin, in contrast to free diffusion, from reductase to luciferase by all enzyme couples tested. Kinetic mechanisms were determined for the FRG(F)-L(F) and FRP(H)-L(F) coupled reactions. For these two and the FRG(F)-L(H) coupled reactions, patterns of FMN inhibition and effects of replacement of the FMN cofactor of FRP(H) and FRG(F) by 2-thioFMN were also characterized. Similar to the FRP(H)-L(H) couple, direct cofactor transfer was detected for FRG(F)-L(F) and FRP(H)-L(F). In contrast, despite the structural similarities between FRG(F) and FRP(H) and between L(F) and L(H), direct flavin product transfer was observed for the FRG(F)-L(H) couple. The mechanism of reduced flavin transfer appears to be delicately controlled by both flavin reductase and luciferase in the couple rather than unilaterally by either enzyme species.  相似文献   

6.
Apoenzyme of the major NAD(P)H-utilizing flavin reductase FRG/FRase I from Vibrio fischeri was prepared. The apoenzyme bound one FMN cofactor per enzyme monomer to yield fully active holoenzyme. The FMN cofactor binding resulted in substantial quenching of both the flavin and the protein fluorescence intensities without any significant shifts in the emission peaks. In addition to FMN binding (K(d) 0.5 microM at 23 degrees C), the apoenzyme also bound 2-thioFMN, FAD and riboflavin as a cofactor with K(d) values of 1, 12, and 37 microM, respectively, at 23 degrees C. The 2-thioFMN containing holoenzyme was about 40% active in specific activity as compared to the FMN-containing holoenzyme. The FAD- and riboflavin-reconstituted holoenzymes were also catalytically active but their specific activities were not determined. FRG/FRase I followed a ping-pong kinetic mechanism. It is proposed that the enzyme-bound FMN cofactor shuttles between the oxidized and the reduced form during catalysis. For both the FMN- and 2-thioFMN-containing holoenzymes, 2-thioFMN was about 30% active as compared to FMN as a substrate. FAD and riboflavin were also active substrates. FRG/FRase I was shown by ultracentrifugation at 4 degrees C to undergo a monomer-dimer equilibrium, with K(d) values of 18.0 and 13.4 microM for the apo- and holoenzymes, respectively. All the spectral, ligand equilibrium binding, and kinetic properties described above are most likely associated with the monomeric species of FRG/FRase I. Many aspects of these properties are compared with a structurally and functionally related Vibrio harveyi NADPH-specific flavin reductase FRP.  相似文献   

7.
B Lei  M Liu  S Huang    S C Tu 《Journal of bacteriology》1994,176(12):3552-3558
NAD(P)H-flavin oxidoreductases (flavin reductases) from luminous bacteria catalyze the reduction of flavin by NAD(P)H and are believed to provide the reduced form of flavin mononucleotide (FMN) for luciferase in the bioluminescence reaction. By using an oligonucleotide probe based on the partial N-terminal amino acid sequence of the Vibrio harveyi NADPH-FMN oxidoreductase (flavin reductase P), a recombinant plasmid, pFRP1, was obtained which contained the frp gene encoding this enzyme. The DNA sequence of the frp gene was determined; the deduced amino acid sequence for flavin reductase P consists of 240 amino acid residues with a molecular weight of 26,312. The frp gene was overexpressed, apparently through induction, in Escherichia coli JM109 cells harboring pFRP1. The cloned flavin reductase P was purified to homogeneity by following a new and simple procedure involving FMN-agarose chromatography as a key step. The same chromatography material was also highly effective in concentrating diluted flavin reductase P. The purified enzyme is a monomer and is unusual in having a tightly bound FMN cofactor. Distinct from the free FMN, the bound FMN cofactor showed a diminished A375 peak and a slightly increased 8-nm red-shifted A453 peak and was completely or nearly nonfluorescent. The Kms for FMN and NADPH and the turnover number of this flavin reductase were determined. In comparison with other flavin reductases and homologous proteins, this flavin reductase P shows a number of distinct features with respect to primary sequence, redox center, and/or kinetic mechanism.  相似文献   

8.
The crystal structure of a major oxygen-insensitive nitroreductase (NfsA) from Escherichia coli has been solved by the molecular replacement method at 1.7-A resolution. This enzyme is a homodimeric flavoprotein with one FMN cofactor per monomer and catalyzes reduction of nitrocompounds using NADPH. The structure exhibits an alpha + beta-fold, and is comprised of a central domain and an excursion domain. The overall structure of NfsA is similar to the NADPH-dependent flavin reductase of Vibrio harveyi, despite definite difference in the spatial arrangement of residues around the putative substrate-binding site. On the basis of the crystal structure of NfsA and its alignment with the V. harveyi flavin reductase and the NADPH-dependent nitro/flavin reductase of Bacillus subtilis, residues Arg(203) and Arg(208) of the loop region between helices I and J in the vicinity of the catalytic center FMN is predicted as a determinant for NADPH binding. The R203A mutant results in a 33-fold increase in the K(m) value for NADPH indicating that the side chain of Arg(203) plays a key role in binding NADPH possibly to interact with the 2'-phosphate group.  相似文献   

9.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

10.
Li X  Chow DC  Tu SC 《Biochemistry》2006,45(49):14781-14787
The Vibrio harveyi NADPH-specific flavin reductase FRP follows a ping-pong mechanism but switches to a sequential mechanism in the luciferase-coupled reaction. The bound FMN co-isolated with FRP, while acting as a genuine cofactor in the single-enzyme reaction, functions in the luciferase-coupled reaction as a prebound substrate and is directly transferred to luciferase once it is reduced [Lei, B., and Tu, S.-C. (1998) Biochemistry 37, 14623-14629]. With the aim of better understanding the functions of FMN in the FRP holoenzyme, this study was undertaken to quantify and compare the thermodynamic properties of the binding of oxidized and reduced FMN by the FRP apoenzyme. By isothermal titration calorimetry (ITC) measurements in various buffers at pH 7.0 and 15-30 degrees C, the binding of FMN by apo-FRP was found to be noncooperative, exothermic, and primarily enthalpy driven. The binding free energy change (hence, the association constant) was nearly invariant over this temperature range. Significant conformational changes in FRP upon binding of FMN were indicated. Equilibrium bindings of reduced flavins by flavin-dependent proteins have rarely been studied. In this work, the thermodynamic properties of binding of reduced FMN by apo-FRP were found to closely resemble those of FMN binding under three sets of experimental conditions via ITC measurements and, in one case, fluorescence quenching. The kinetically deduced ping-pong mechanism of FRP is now supported by direct measurements of binding affinities of the oxidized and reduced FMN cofactors. These findings are also discussed in relation to the function of FRP as a reduced flavin donor in the FRP-luciferase couple.  相似文献   

11.
The alkanesulfonate monooxygenase system from Escherichia coli is involved in scavenging sulfur from alkanesulfonates under sulfur starvation. An FMN reductase (SsuE) catalyzes the reduction of FMN by NADPH, and the reduced flavin is transferred to the monooxygenase (SsuD). Rapid reaction kinetic analyses were performed to define the microscopic steps involved in SsuE catalyzed flavin reduction. Results from single-wavelength analyses at 450 and 550 nm showed that reduction of FMN occurs in three distinct phases. Following a possible rapid equilibrium binding of FMN and NADPH to SsuE (MC-1) that occurs before the first detectable step, an initial fast phase (241 s(-1)) corresponds to the interaction of NADPH with FMN (CT-1). The second phase is a slow conversion (11 s(-1)) to form a charge-transfer complex of reduced FMNH(2) with NADP(+) (CT-2), and represents electron transfer from the pyridine nucleotide to the flavin. The third step (19 s(-1)) is the decay of the charge-transfer complex to SsuE with bound products (MC-2) or product release from the CT-2 complex. Results from isotope studies with [(4R)-(2)H]NADPH demonstrates a rate-limiting step in electron transfer from NADPH to FMN, and may imply a partial rate-limiting step from CT-2 to MC-2 or the direct release of products from CT-2. While the utilization of flavin as a substrate by the alkanesulfonate monooxygenase system is novel, the mechanism for flavin reduction follows an analogous reaction path as standard flavoproteins.  相似文献   

12.
Chung HW  Tu SC 《Biochemistry》2012,51(24):4880-4887
Vibrio harveyi NADPH-FMN oxidoreductase (FRP) catalyzes flavin reduction by NADPH. In comparing amino acid sequence and crystal structure with Escherichia coli NfsA, residues N134, R225, R133, K167, and R15 were targeted for investigation of their possible roles in the binding and utilization of the NADPH substrate. By mutation of each of these five residues to an alanine, steady-state rate analyses showed that the variants K167A and R15A had apparently greatly increased K(m,NADPH) and reduced k(cat)/K(m,NADPH), whereas little or much more modest changes were found for the other variants. The deuterium isotope effects (D)(V/K) for (4R)-[4-(2)H]-NADPH were markedly increased to 6.3 and 7.4 for K167A and R15A, respectively, indicating that the rate constants for NADPH and NADP(+) dissociation were greatly enhanced relative to the hydride transfer steps. Also, anaerobic stopped-flow analyses revealed that the equilibrium dissociation constant for NADPH binding (K(d)) to be 2.5-3.9 and 1.1 mM for K167A and R15A, respectively, much higher than the 0.4 μM K(d) for the native FRP, whereas the k(cat) of these two variants were similar to that of the wild-type enzyme. Moreover, the K167 to alanine mutation led to even a slight increase in k(cat)/K(m) for NADH. These results, taken together, provide a strong support to the conclusion that K167 and R15 each was critical in the binding of NADPH by FRP. Such a functional role may also exist for other FRP homologous proteins.  相似文献   

13.
Garnaud PE  Koetsier M  Ost TW  Daff S 《Biochemistry》2004,43(34):11035-11044
Electron transfer through neuronal nitric oxide synthase (nNOS) is regulated by the reversible binding of calmodulin (CaM) to the reductase domain of the enzyme, the conformation of which has been shown to be dependent on the presence of substrate, NADPH. Here we report the preparation of the isolated flavin mononucleotide (FMN)-binding domain of nNOS with bound CaM and the electrochemical analysis of this and the isolated flavin adenine dinucleotide (FAD)-binding domain in the presence and absence of NADP(+) and ADP (an inhibitor). The FMN-binding domain was found to be stable only in the presence of bound CaM/Ca(2+), removal of which resulted in precipitation of the protein. The FMN formed a kinetically stabilized blue semiquinone with an oxidized/semiquinone reduction potential of -179 mV. This is 80 mV more negative than the potential of the FMN in the isolated reductase domain, that is, in the presence of the FAD-binding domain. The FMN semiquinone/hydroquinone redox couple was found to be similar in both constructs. The isolated FAD-binding domain, generated by controlled proteolysis of the reductase domain, was found to have similar FAD reduction potentials to the isolated reductase domain. Both formed a FAD-hydroquinone/NADP(+) charge-transfer complex with a long-wavelength absorption band centered at 780 nm. Formation of this complex resulted in thermodynamic destabilization of the FAD semiquinone relative to the hydroquinone and a 30 mV increase in the FAD semiquinone/hydroquinone reduction potential. Binding of ADP, however, had little effect. The possible role of the nicotinamide/FADH(2) stacking interaction in controlling electron transfer and its likely dependence on protein conformation are discussed.  相似文献   

14.
Phe(1395) stacks parallel to the FAD isoalloxazine ring in neuronal nitric-oxide synthase (nNOS) and is representative of conserved aromatic amino acids found in structurally related flavoproteins. This laboratory previously showed that Phe(1395) was required to obtain the electron transfer properties and calmodulin (CaM) response normally observed in wild-type nNOS. Here we characterized the F1395S mutant of the nNOS flavoprotein domain (nNOSr) regarding its physical properties, NADP(+) binding characteristics, flavin reduction kinetics, steady-state and pre-steady-state cytochrome c reduction kinetics, and ability to shield its FMN cofactor in response to CaM or NADP(H) binding. F1395S nNOSr bound NADP(+) with 65% more of the nicotinamide ring in a productive conformation with FAD for hydride transfer and had an 8-fold slower rate of NADP(+) dissociation. CaM stimulated the rates of NADPH-dependent flavin reduction in wild-type nNOSr but not in the F1395S mutant, which had flavin reduction kinetics similar to those of CaM-free wild-type nNOSr. CaM-free F1395S nNOSr lacked repression of cytochrome c reductase activity that is typically observed in nNOSr. The combined results from pre-steady-state and EPR experiments revealed that this was associated with a lesser degree of FMN shielding in the NADP(+)-bound state as compared with wild type. We conclude that Phe(1395) regulates nNOSr catalysis in two ways. It facilitates NADP(+) release to prevent this step from being rate-limiting, and it enables NADP(H) to properly regulate a conformational equilibrium involving the FMN subdomain that controls reactivity of the FMN cofactor in electron transfer.  相似文献   

15.
Russell TR  Demeler B  Tu SC 《Biochemistry》2004,43(6):1580-1590
The homodimeric NADH:flavin oxidoreductase from Aminobacter aminovorans is an NADH-specific flavin reductase herein designated FRD(Aa). FRD(Aa) was characterized with respect to purification yields, thermal stability, isoelectric point, molar absorption coefficient, and effects of phosphate buffer strength and pH on activity. Evidence from this work favors the classification of FRD(Aa) as a flavin cofactor-utilizing class I flavin reductase. The isolated native FRD(Aa) contained about 0.5 bound riboflavin-5'-phosphate (FMN) per enzyme monomer, but one bound flavin cofactor per monomer was obtainable in the presence of excess FMN or riboflavin. In addition, FRD(Aa) holoenzyme also utilized FMN, riboflavin, or FAD as a substrate. Steady-state kinetic results of substrate titrations, dead-end inhibition by AMP and lumichrome, and product inhibition by NAD(+) indicated an ordered sequential mechanism with NADH as the first binding substrate and reduced FMN as the first leaving product. This is contrary to the ping-pong mechanism shown by other class I flavin reductases. The FMN bound to the native FRD(Aa) can be fully reduced by NADH and subsequently reoxidized by oxygen. No NADH binding was detected using 90 microM FRD(Aa) apoenzyme and 300 microM NADH. All results favor the interpretation that the bound FMN was a cofactor rather than a substrate. It is highly unusual that a flavin reductase using a sequential mechanism would require a flavin cofactor to facilitate redox exchange between NADH and a flavin substrate. FRD(Aa) exhibited a monomer-dimer equilibrium with a K(d) of 2.7 microM. Similarities and differences between FRD(Aa) and certain flavin reductases are discussed.  相似文献   

16.
The neuronal nitric-oxide synthase (nNOS) flavoprotein domain (nNOSr) contains regulatory elements that repress its electron flux in the absence of bound calmodulin (CaM). The repression also requires bound NADP(H), but the mechanism is unclear. The crystal structure of a CaM-free nNOSr revealed an ionic interaction between Arg(1400) in the C-terminal tail regulatory element and the 2'-phosphate group of bound NADP(H). We tested the role of this interaction by substituting Ser and Glu for Arg(1400) in nNOSr and in the full-length nNOS enzyme. The CaM-free nNOSr mutants had cytochrome c reductase activities that were less repressed than in wild-type, and this effect could be mimicked in wild-type by using NADH instead of NADPH. The nNOSr mutants also had faster flavin reduction rates, greater apparent K(m) for NADPH, and greater rates of flavin auto-oxidation. Single-turnover cytochrome c reduction data linked these properties to an inability of NADP(H) to cause shielding of the FMN module in the CaM-free nNOSr mutants. The full-length nNOS mutants had no NO synthesis in the CaM-free state and had lower steady-state NO synthesis activities in the CaM-bound state compared with wild-type. However, the mutants had faster rates of ferric heme reduction and ferrous heme-NO complex formation. Slowing down heme reduction in R1400E nNOS with CaM analogues brought its NO synthesis activity back up to normal level. Our studies indicate that the Arg(1400)-2'-phosphate interaction is a means by which bound NADP(H) represses electron transfer into and out of CaM-free nNOSr. This interaction enables the C-terminal tail to regulate a conformational equilibrium of the FMN module that controls its electron transfer reactions in both the CaM-free and CaM-bound forms of nNOS.  相似文献   

17.
Bright luminescent yeast cells with light intensities similar to bacteria containing luciferase (LuxAB) were generated by providing saturating nontoxic levels of the substrates for the bioluminescence reaction (FMNH(2)+O(2) and fatty aldehyde-->light). Z-9-Tetradecenal added to yeast (+luxAB) gave a luminescent signal close to that with decanal with the signal remaining strong for >24h while luminescence of yeast with decanal decayed to less than 0.01% of that with Z-9-tetradecenal after 2min. Moreover, yeast survived in 0.5% (v/v) Z-9-tetradecenal while 0.005% (v/v) decanal was lethal. Luminescence of yeast (+luxAB) was also stimulated 100-fold by transformation with the NADPH-specific FMN reductase (FRP) from Vibrio harveyi. The recognition of the nontoxicity and high luminescence generated by Z-9-tetradecenal and the generation of high levels of FMNH(2) in yeast by transformation with a flavin reductase provide evidence for the strong potential use of bacterial luciferase as the light-emitting sensor of choice in eukaryotic organisms.  相似文献   

18.
The crystal structure of NADPH-cytochrome P450 reductase (CYPOR) implies that a large domain movement is essential for electron transfer from NADPH via FAD and FMN to its redox partners. To test this hypothesis, a disulfide bond was engineered between residues Asp(147) and Arg(514) in the FMN and FAD domains, respectively. The cross-linked form of this mutant protein, designated 147CC514, exhibited a significant decrease in the rate of interflavin electron transfer and large (≥90%) decreases in rates of electron transfer to its redox partners, cytochrome c and cytochrome P450 2B4. Reduction of the disulfide bond restored the ability of the mutant to reduce its redox partners, demonstrating that a conformational change is essential for CYPOR function. The crystal structures of the mutant without and with NADP(+) revealed that the two flavin domains are joined by a disulfide linkage and that the relative orientations of the two flavin rings are twisted ~20° compared with the wild type, decreasing the surface contact area between the two flavin rings. Comparison of the structures without and with NADP(+) shows movement of the Gly(631)-Asn(635) loop. In the NADP(+)-free structure, the loop adopts a conformation that sterically hinders NADP(H) binding. The structure with NADP(+) shows movement of the Gly(631)-Asn(635) loop to a position that permits NADP(H) binding. Furthermore, comparison of these mutant and wild type structures strongly suggests that the Gly(631)-Asn(635) loop movement controls NADPH binding and NADP(+) release; this loop movement in turn facilitates the flavin domain movement, allowing electron transfer from FMN to the CYPOR redox partners.  相似文献   

19.
Human gene for pancreas protein 2 (PAN2) is a novel member of the short-chain dehydrogenase/reductase gene superfamily. The properties of PAN2 protein have not yet been characterized. We present the first evidence that human PAN2 is a ubiquitously expressed microsomal enzyme that recognizes retinoids but not steroids as substrates with the apparent K(m) values between 0.08 microM and 0.4 microM. PAN2 is approximately 4-fold more efficient in the reductive than in the oxidative direction. The apparent K(m) values for NADP(+) and NADPH are 0.65 microM and 0.32 microM versus 1200 microM and 1060 microM for NAD(+) and NADH, respectively. Kinetic constants and expression pattern of PAN2 suggest that it is likely to function as a reductase in vivo and might contribute to the reduction of retinaldehyde to retinol in most human tissues.  相似文献   

20.
S Zenno  K Saigo 《Journal of bacteriology》1994,176(12):3544-3551
Genes encoding NAD(P)H-flavin oxidoreductases (flavin reductases) similar in both size and sequence to Fre, the most abundant flavin reductase in Escherichia coli, were identified in four species of luminous bacteria, Photorhabdus luminescens (ATCC 29999), Vibrio fischeri (ATCC 7744), Vibrio harveyi (ATCC 33843), and Vibrio orientalis (ATCC 33934). Nucleotide sequence analysis showed Fre-like flavin reductases in P. luminescens and V. fischeri to consist of 233 and 236 amino acids, respectively. As in E. coli Fre, Fre-like enzymes in luminous bacteria preferably used riboflavin as an electron acceptor when NADPH was used as an electron donor. These enzymes also were good suppliers of reduced flavin mononucleotide (FMNH2) to the bioluminescence reaction. In V. fischeri, the Fre-like enzyme is a minor flavin reductase representing < 10% of the total FMN reductase. That the V. fischeri Fre-like enzyme has no appreciable homology in amino acid sequence to the major flavin reductase in V. fischeri, FRase I, indicates that at least two different types of flavin reductases supply FMNH2 to the luminescence system in V. fischeri. Although Fre-like flavin reductases are highly similar in sequence to luxG gene products (LuxGs), Fre-like flavin reductases and LuxGs appear to constitute two separate groups of flavin-associated proteins.  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