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.
R67 dihydrofolate reductase (DHFR) catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate using NADPH as a cofactor. This enzyme is a homotetramer possessing 222 symmetry, and a single active site pore traverses the length of the protein. A promiscuous binding surface can accommodate either DHF or NADPH, thus two nonproductive complexes can form (2NADPH or 2DHF) as well as a productive complex (NADPH.DHF). The role of water in binding was monitored using a number of different osmolytes. From isothermal titration calorimetry (ITC) studies, binding of NADPH is accompanied by the net release of 38 water molecules. In contrast, from both steady state kinetics and ITC studies, binding of DHF is accompanied by the net uptake of water. Although different osmolytes have similar effects on NADPH binding, variable results are observed when DHF binding is probed. Sensitivity to water activity can also be probed by an in vivo selection using the antibacterial drug, trimethoprim, where the water content of the media is decreased by increasing concentrations of sorbitol. The ability of wild type and mutant clones of R67 DHFR to allow host Escherichia coli to grow in the presence of trimethoprim plus added sorbitol parallels the catalytic efficiency of the DHFR clones, indicating water content strongly correlates with the in vivo function of R67 DHFR.  相似文献   

2.
Duff MR  Grubbs J  Serpersu E  Howell EE 《Biochemistry》2012,51(11):2309-2318
Previous osmotic stress studies on the role of solvent in two structurally unrelated dihydrofolate reductases (DHFRs) found weaker binding of dihydrofolate (DHF) to either enzyme in the presence of osmolytes. To explain these unusual results, weak interactions between DHF and osmolytes were proposed, with a competition between osmolyte and DHFR for DHF. High osmolyte concentrations will inhibit binding of the cognate pair. To evaluate this hypothesis, we devised a small molecule approach. Dimerization of folate, monitored by nuclear magnetic resonance, was weakened 2-3-fold upon addition of betaine or dimethyl sulfoxide (DMSO), supporting preferential interaction of either osmolyte with the monomer (as it possesses a larger surface area). Nuclear Overhauser effect (NOE) spectroscopy experiments found a positive NOE for the interaction of the C3'/C5' benzoyl ring protons with the C9 proton in buffer; however, a negative NOE was observed upon addition of betaine or DMSO. This change indicated a decreased tumbling rate, consistent with osmolyte interaction. Osmotic stress experiments also showed that betaine, DMSO, and sucrose preferentially interact with folate. Further, studies with the folate fragments, p-aminobenzoic acid and pterin 6-carboxylate, revealed interactions for both model compounds with betaine and sucrose. In contrast, DMSO was strongly excluded from the pterin ring but preferentially interacted with the p-aminobenzoyl moiety. These interactions are likely to be important in vivo because of the crowded conditions of the cell where weak contacts can more readily compete with specific binding interactions.  相似文献   

3.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

4.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  相似文献   

5.
Chopra S  Lynch R  Kim SH  Jackson M  Howell EE 《Biochemistry》2006,45(21):6596-6605
R67 dihydrofolate reductase (DHFR) is a novel homotetrameric protein that possesses 222 symmetry and a single, voluminous active site pore. This symmetry poses numerous limitations on catalysis; for example, two dihydrofolate (DHF) molecules or two NADPH molecules, or one substrate plus one cofactor can bind. Only the latter combination leads to catalysis. To garner additional information on how this enzyme facilitates transition-state formation, the temperature dependence of binding and catalysis was monitored. The binding of NADPH and DHF is enthalpy-driven. Previous primary isotope effect studies indicate hydride transfer is at least partially rate-determining. Accordingly, the activation energy associated with transition-state formation was measured and is found to be 6.9 kcal/mol (DeltaH(++)(25) = 6.3 kcal/mol). A large entropic component is also found associated with catalysis, TDeltaS(++)(25) = -11.3 kcal/mol. The poor substrate, dihydropteroate, binds more weakly than dihydrofolate (DeltaDeltaG = 1.4 kcal/mol) and displays a large loss in the binding enthalpy value (DeltaDeltaH = 3.8 kcal/mol). The k(cat) value for dihydropteroate reduction is decreased 1600-fold compared to DHF usage. This effect appears to derive mostly from the DeltaDeltaH difference in binding, demonstrating that the glutamate tail is important for catalysis. This result is surprising, as the para-aminobenzoyl-glutamate tail of DHF has been previously shown to be disordered by both NMR and crystallography studies. Viscosity studies were also performed and confirmed that the hydride transfer rate is not sensitive to sucrose addition. Surprisingly, binding of DHF, by both K(m) and K(d) determination, was found to be sensitive to added viscogens, suggesting a role for water in DHF binding.  相似文献   

6.
The interaction of type II R67 dihydrofolate reductase (DHFR) with its cofactor nicotinamide adenine dinucleotide phosphate (NADP(+)) has been studied using nuclear magnetic resonance (NMR). Doubly labeled [U-(13)C,(15)N]DHFR was obtained from Escherichia coli grown on a medium containing [U-(13)C]-D-glucose and (15)NH(4)Cl, and the 16 disordered N-terminal amino acids were removed by treatment with chymotrypsin. Backbone and side chain NMR assignments were made using triple-resonance experiments. The degeneracy of the amide (1)H and (15)N shifts of the tetrameric DHFR was preserved upon addition of NADP(+), consistent with kinetic averaging among equivalent binding sites. Analysis of the more titration-sensitive DHFR amide resonances as a function of added NADP(+) gave a K(D) of 131 +/- 50 microM, consistent with previous determinations using other methodology. We have found that the (1)H spectrum of NADP(+) in the presence of the R67 DHFR changes as a function of time. Comparison with standard samples and mass spectrometric analysis indicates a slow conversion of NADP(+) to NAD(+), i.e., an apparent NADP(+) phosphatase activity. Studies of this activity in the presence of folate and a folate analogue support the conclusion that this activity results from an interaction with the DHFR rather than a contaminating phosphatase. (1)H NMR studies of a mixture of NADP(+) and NADPH in the presence of the enzyme reveal that a ternary complex forms in which the N-4A and N-4B nuclei of the NADPH are in the proximity of the N-4 and N-5 nuclei of NADP(+). Studies using the NADP(+) analogue acetylpyridine adenosine dinucleotide phosphate (APADP(+)) demonstrated a low level of enzyme-catalyzed hydride transfer from NADPH. Analysis of DHFR backbone dynamics revealed little change upon binding of NADP(+). These additional catalytic activities and dynamic behavior are in marked contrast to those of type I DHFR.  相似文献   

7.
Individual flowers from inflorescences of Dahlia variabilis (cv Scarlet Star) in young developmental stages contained relatively high activity of (+)-dihydroflavonol (DHF) 4-reductase. The DHF reductase was purified from such flowers to apparent homogeneity by a five-step procedure. This included affinity adsorption on Blue Sepharose and elution of the enzyme with NADP+. By gel filtration and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis it was shown that DHF reductase contains only one polypeptide chain with a Mr of about 41,000. The reductase required NADPH as cofactor and catalyzed transfer of the pro-S hydrogen of NADPH to the substrate. Flavanones and dihydroflavonols (3-hydroxyflavanones) were substrates for DHF reductase with pH optima of about 6.0 for flavanones and of about 6.8 for dihydroflavonols. Flavanones were reduced to the corresponding flavan-4-ols and (+)-dihydroflavonols to flavan-3,4-cis-diols. Apparent Michaelis constants determined for (2S)-naringenin, (2S)-eriodicytol, (+)-dihydrokaempferol, (+)-dihydroquercetin, and NADPH were, respectively, 2.3, 2, 10, 15, and 42 microM. V/Km values were higher for dihydroflavonols than for flavanones. Conversion of dihydromyricetin to leucodelphinidin was also catalyzed by the enzyme at a low rate, whereas flavones and flavonols were not accepted as substrates. DHF reductase was not inhibited by metal chelators.  相似文献   

8.
Smiley RD  Stinnett LG  Saxton AM  Howell EE 《Biochemistry》2002,41(52):15664-15675
R67 dihydrofolate reductase (DHFR) is an enzyme, encoded by an R-plasmid, that confers resistance to the antibacterial agent trimethoprim. This homotetramer possesses a single active site pore and exact 222 symmetry. The symmetry imposes constraints on the ability of the enzyme to optimize binding of the substrate, dihydrofolate (DHF), and the cofactor, NADPH, resulting in a "one site fits both ligands" approach. This approach allows formation of either a NADPH.NADPH, dihydrofolate.dihydrofolate, or NADPH.dihydrofolate complex. The first two complexes are nonproductive, while the third is the productive catalytic species. To break the symmetry of the active site, a tandem array of four R67 DHFR genes has been linked in frame, allowing individual manipulation of each gene copy. Various numbers and combinations of asymmetric Q67H mutations have been engineered into the tandem gene array. The Q67H mutation was chosen for investigation as it was previously found to tighten binding to both dihydrofolate and NADPH by approximately 100-fold in homotetrameric R67 DHFR [Park, H., Bradrick, T. D., and Howell, E. E. (1997) Protein Eng. 10, 1415-1424]. Nonadditive effects on ligand binding are observed when one to four mutations are inserted, indicating either conformational changes in the protein or different cooperativity patterns in the ligand-ligand interactions. From steady state kinetics, addition of Q67H mutations does not drastically affect formation of the NADPH.dihydrofolate complex; however, a large energy difference between the productive and nonproductive complexes is no longer maintained. A role for Q67 in discriminating between these various states is proposed. Since theories of protein evolution suggest gene duplication followed by accumulation of mutations can lead to divergence of activity, this study is a first step toward asking if introduction of asymmetric mutations in the quadrupled R67 DHFR gene can lead to optimization of ligand binding sites.  相似文献   

9.
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   

10.
Krahn JM  Jackson MR  DeRose EF  Howell EE  London RE 《Biochemistry》2007,46(51):14878-14888
Type II dihydrofolate reductase (DHFR) is a plasmid-encoded enzyme that confers resistance to bacterial DHFR-targeted antifolate drugs. It forms a symmetric homotetramer with a central pore which functions as the active site. Its unusual structure, which results in a promiscuous binding surface that accommodates either the dihydrofolate (DHF) substrate or the NADPH cofactor, has constituted a significant limitation to efforts to understand its substrate specificity and reaction mechanism. We describe here the first structure of a ternary R67 DHFR.DHF.NADP+ catalytic complex, resolved to 1.26 A. This structure provides the first clear picture of how this enzyme, which lacks the active site carboxyl residue that is ubiquitous in Type I DHFRs, is able to function. In the catalytic complex, the polar backbone atoms of two symmetry-related I68 residues provide recognition motifs that interact with the carboxamide on the nicotinamide ring, and the N3-O4 amide function on the pteridine ring. This set of interactions orients the aromatic rings of substrate and cofactor in a relative endo geometry in which the reactive centers are held in close proximity. Additionally, a central, hydrogen-bonded network consisting of two pairs of Y69-Q67-Q67'-Y69' residues provides an unusually tight interface, which appears to serve as a "molecular clamp" holding the substrates in place in an orientation conducive to hydride transfer. In addition to providing the first clear insight regarding how this extremely unusual enzyme is able to function, the structure of the ternary complex provides general insights into how a mutationally challenged enzyme, i.e., an enzyme whose evolution is restricted to four-residues-at-a-time active site mutations, overcomes this fundamental limitation.  相似文献   

11.
Feng J  Goswami S  Howell EE 《Biochemistry》2008,47(2):555-565
R67 dihydrofolate reductase (DHFR) bears no sequence or structural homologies with chromosomal DHFRs. The gene for this enzyme produces subunits that are 78 amino acids long, which assemble into a homotetramer possessing 222 symmetry. More recently, a tandem array of four gene copies linked in-frame was constructed, which produces a monomer containing 312 amino acids named Quad3. Asymmetric mutations in Quad3 have also been constructed to probe the role of Q67 and K32 residues in catalysis. This present study mixes and matches mutations to determine if the Q67H mutation, which tightens binding approximately 100-fold to both dihydrofolate (DHF) and NADPH, can help rescue the K32M mutation. While the latter mutation weakens DHF binding over 60-fold, it concurrently increases kcat by a factor of 5. Two Q67H mutations were added to gene copies 1 and 4 in conjunction with the K32M mutation in gene copies 1 and 3. Addition of these Q67H mutations tightens binding 40-fold, and the catalytic efficiency (kcat/Km(DHF)) of the resulting protein is similar to that of Quad3. Since these Q67H mutations can mostly compensate for the K32M lesion, K32 must not be necessary for DHF binding. Another multimutant combines the K32M mutation in gene copies 1 and 3 with the Q67H mutation in all gene copies. This mutant is inhibited by DHF but not NADPH, indicating that NADPH binds only to the wild type half of the pore, while DHF can bind to either the wild type or mutant half of the pore. This inhibition pattern contrasts with the mutant containing only the Q67H substitution in all four gene copies, which is severely inhibited by both NADPH and substrate. Since gene duplication and divergence are evolutionary tools for gaining function, these constructs are a first step toward building preferences for NADPH and DHF in each half of the active site pore of this primitive enzyme.  相似文献   

12.
Heteronuclear NMR methods have been used to probe the conformation of four complexes of Escherichia coli dihydrofolate reductase (DHFR) in solution. (1)H(N), (15)N, and (13)C(alpha) resonance assignments have been made for the ternary complex with folate and oxidized NADP(+) cofactor and the ternary complex with folate and a reduced cofactor analog, 5,6-dihydroNADPH. The backbone chemical shifts have been compared with those of the binary complex of DHFR with the substrate analog folate and the binary complex with NADPH (the holoenzyme). Analysis of (1)H(N) and (15)N chemical shifts has led to the identification of marker resonances that report on the active site conformation of the enzyme. Other backbone amide resonances report on the presence of ligands in the pterin binding pocket and in the adenosine and nicotinamide-ribose binding sites of the NADPH cofactor. The chemical shift data indicate that the enzyme populates two dominant structural states in solution, with the active site loops in either the closed or occluded conformations defined by X-ray crystallography; there is no evidence that the open conformation observed in some X-ray structures of E. coli DHFR are populated in solution.  相似文献   

13.
The ferredoxin nicotinamide adenine dinucleotide phosphate reductase from Pseudomonas aeruginosa ( pa-FPR) in complex with NADP (+) has been characterized by X-ray crystallography and in solution by NMR spectroscopy. The structure of the complex revealed that pa-FPR harbors a preformed NADP (+) binding pocket where the cofactor binds with minimal structural perturbation of the enzyme. These findings were complemented by obtaining sequential backbone resonance assignments of this 29518 kDa enzyme, which enabled the study of the pa-FPR-NADP complex by monitoring chemical shift perturbations induced by addition of NADP (+) or the inhibitor adenine dinucleotide phosphate (ADP) to pa-FPR. The results are consistent with a preformed NADP (+) binding site and also demonstrate that the pa-FPR-NADP complex is largely stabilized by interactions between the protein and the 2'-P AMP portion of the cofactor. Analysis of the crystal structure also shows a vast network of interactions between the two cofactors, FAD and NADP (+), and the characteristic AFVEK (258) C'-terminal extension that is typical of bacterial FPRs but is absent in their plastidic ferredoxin NADP (+) reductase (FNR) counterparts. The conformations of NADP (+) and FAD in pa-FPR place their respective nicotinamide and isoalloxazine rings 15 A apart and separated by residues in the C'-terminal extension. The network of interactions among NADP (+), FAD, and residues in the C'-terminal extension indicate that the gross conformational rearrangement that would be necessary to place the nicotinamide and isoalloxazine rings parallel and adjacent to one another for direct hydride transfer between NADPH and FAD in pa-FPR is highly unlikely. This conclusion is supported by observations made in the NMR spectra of pa-FPR and the pa-FPR-NADP complex, which strongly suggest that residues in the C'-terminal sequence do not undergo conformational exchange in the presence or absence of NADP (+). These findings are discussed in the context of a possible stepwise electron-proton-electron transfer of hydride in the oxidation of NADPH by FPR enzymes.  相似文献   

14.
Circular dichroism has been used to monitor the binding of pyridine nucleotide cofactors to enzyme-folate analog complexes of dihydrofolate reductase from Escherichia coli B (MB 1428). The enzyme binds one molar equivalent of many folate analogs and two molar equivalents of several pyridine nucleotide cofactors. The apo-enzyme has very low optical activity. The binding of folate analogs including folate, dihydrofolate, methotrexate, trimethoprim and pyrimethamine induce large Cotton effects. Pyridine nucleotides when bound to the enzyme-folate analog complexes also induce new optically active bands; all the effects being due to the first molar equivalent of cofactor bound. NADPH and NADP+ induce very similar bands when bound to the enzyme-methotrexate complex suggesting that the geometry of the complexes formed are very similar. The oxidized and reduced cofactor likewise have similar effects on the enzyme-folate complex. However, NADPH and NADP+ addition to both the enzyme-trimethoprim and enzyme-pyrimethamine complexes have significantly different effects on the circular dichroism spectra, suggesting that the inhibitors which are less homologous to the natural dihydrofolate substrate allow more conformational freedom in the enzyme-inhibitor-cofactor complex. In most cases the prior binding of the folate analog greatly increases the binding of the first molar equivalent of cofactor so that at concentrations of approx. 5-20 muM the binding appears stoichiometric. Pyrimethamine is an exception in that it apparently has no effect on the binding of NADPH to the enzyme.  相似文献   

15.
D J Murphy  S J Benkovic 《Biochemistry》1989,28(7):3025-3031
The strictly conserved residue leucine-54 of Escherichia coli dihydrofolate reductase forms part of the hydrophobic wall which binds the p-aminobenzoyl side chain of dihydrofolate. In addition to the previously reported glycine-54 mutant, isoleucine-54 and asparagine-54 substitutions have been constructed and characterized with regard to their effects on binding and catalysis. NADP+ and NADPH binding is virtually unaffected with the exception of a 15-fold decrease in NADPH dissociation from the Gly-54 mutant. The synergistic effect of NADPH on tetrahydrofolate dissociation seen in the wild-type enzyme is lost in the isoleucine-54 mutant: little acceleration is seen in tetrahydrofolate dissociation when cofactor is bound, and there is no discrimination between reduced and oxidized cofactor. The dissociation constants for dihydrofolate and methotrexate increase in the order Leu less than Ile less than Asn less than Gly, varying by a maximum factor of 1700 for dihydrofolate and 6300 for methotrexate. Despite these large changes in binding affinity, the hydride transfer rate of 950 s-1 in the wild-type enzyme is decreased by a constant factor of ca. 30 (2 kcal/mol) regardless of the mutant. Thus, the contributions of residue 54 to binding and catalysis appear to have been separated.  相似文献   

16.
Incubation of malic enzyme (L-malate:NADP+ oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) with ethoxyformic anhydride caused the time-dependent loss of its ability to catalyze reactions requiring the nucleotide cofactor NADP+ or NADPH, such as the oxidative decarboxylase, the NADP+ - stimualted oxalacetate decarboxylase, the pyruvate reductase, and the pyruvate-medium proton exchange activities. Similar loss of oxidative decarboxylase and pyruvate reductase activities was affected by photo-oxidation in the presence of rose bengal. The inactivation of oxidative decarboxylase activity by ethoxyformic anhydride was accompanied by the reaction of greater than or equal to 2.3 histidyl residues per enzyme site and was strongly inhibited by NADP+. Ethoxyformylation also impaired the ability of malic enzyme to bind NADP+ or NADPH. These results support the involvement of histidyl residue(s) at the nucleotide binding site of malic enzyme.  相似文献   

17.
beta-Ketoacyl-acyl carrier protein reductase (FabG) is a key component in the type II fatty acid synthase system. The structures of Escherichia coli FabG and the FabG[Y151F] mutant in binary complexes with NADP(H) reveal that mechanistically important conformational changes accompany cofactor binding. The active site Ser-Tyr-Lys triad is repositioned into a catalytically competent constellation, and a hydrogen bonded network consisting of ribose hydroxyls, the Ser-Tyr-Lys triad, and four water molecules creates a proton wire to replenish the tyrosine proton donated during catalysis. Also, a disordered loop in FabG forms a substructure in the complex that shapes the entrance to the active site. A key observation is that the nicotinamide portion of the cofactor is disordered in the FabG[Y151F].NADP(H) complex, and Tyr151 appears to be necessary for high-affinity cofactor binding. Biochemical data confirm that FabG[Y151F] is defective in NADPH binding. Finally, structural changes consistent with the observed negative cooperativity of FabG are described.  相似文献   

18.
Marohnic CC  Bewley MC  Barber MJ 《Biochemistry》2003,42(38):11170-11182
Microsomal cytochrome b(5) reductase (EC 1.6.2.2) catalyzes the reduction of ferricytochrome b(5) using NADH as the physiological electron donor. Site-directed mutagenesis has been used to engineer the soluble rat cytochrome b(5) reductase diaphorase domain to utilize NADPH as the preferred electron donor. Single and double mutations at residues D239 and F251 were made in a recombinant expression system that corresponded to D239E, S and T, F251R, and Y, D239S/F251R, D239S/F251Y, and D239T/F251R, respectively. Steady-state turnover measurements indicated that D239S/F251Y was bispecific while D239T, D239S/F251R, and D239T/F251R were each NADPH-specific. Wild-type (WT) cytochrome b(5) reductase showed a 3700-fold preference for NADH whereas the mutant with the highest NADPH efficiency, D239T, showed an 11-fold preference for NADPH, a 39200-fold increase. Wild-type cytochrome b(5) reductase only formed a stable charge-transfer complex with NADH while D239T formed complexes with both NADH and NADPH. The rates of hydride ion transfer, determined by stopped-flow kinetics, were k(NADH-WT) = 130 s(-1), k(NADPH-WT) = 5 s(-1), k(NADH-D239T) = 180 s(-1), and k(NADPH-D239T) = 73 s(-1). K(s) determinations by differential spectroscopy demonstrated that D239T could bind nonreducing pyridine nucleotides with a phosphate or a hydroxyl substituent at the 2' position, whereas wild-type cytochrome b(5) reductase would only bind 2' hydroxylated molecules. Oxidation-reduction potentials (E degrees ', n = 2) for the flavin cofactor were WT = -268 mV, D239T = -272 mV, WT+NAD(+) = -190 mV, D239T+NAD(+) = -206 mV, WT+NADP(+) = -253 mV, and D239T+NADP(+) = -215 mV, which demonstrated the thermodynamic contribution of NADP(+) binding to D239T. The crystal structures of D239T and D239T in complex with NAD(+) indicated that the loss of the negative electrostatic surface that precluded 2' phosphate binding in the wild-type enzyme was primarily responsible for the observed improvement in the use of NADPH by the D239T mutant.  相似文献   

19.
Hydroxysteroid dehydrogenases (HSDs) are essential for the biosynthesis and mechanism of action of all steroid hormones. We report the complete kinetic mechanism of a mammalian HSD using rat 3alpha-HSD of the aldo-keto reductase superfamily (AKR1C9) with the substrate pairs androstane-3,17-dione and NADPH (reduction) and androsterone and NADP(+) (oxidation). Steady-state, transient state kinetics, and kinetic isotope effects reconciled the ordered bi-bi mechanism, which contained 9 enzyme forms and permitted the estimation of 16 kinetic constants. In both reactions, loose association of the NADP(H) was followed by two conformational changes, which increased cofactor affinity by >86-fold. For androstane-3,17-dione reduction, the release of NADP(+) controlled k(cat), whereas the chemical event also contributed to this term. k(cat) was insensitive to [(2)H]NADPH, whereas (D)k(cat)/K(m) and the (D)k(lim) (ratio of the maximum rates of single turnover) were 1.06 and 2.06, respectively. Under multiple turnover conditions partial burst kinetics were observed. For androsterone oxidation, the rate of NADPH release dominated k(cat), whereas the rates of the chemical event and the release of androstane-3,17-dione were 50-fold greater. Under multiple turnover conditions full burst kinetics were observed. Although the internal equilibrium constant favored oxidation, the overall K(eq) favored reduction. The kinetic Haldane and free energy diagram confirmed that K(eq) was governed by ligand binding terms that favored the reduction reactants. Thus, HSDs in the aldo-keto reductase superfamily thermodynamically favor ketosteroid reduction.  相似文献   

20.
A C Price  Y M Zhang  C O Rock  S W White 《Biochemistry》2001,40(43):12772-12781
The structure of beta-ketoacyl-[acyl carrier protein] reductase (FabG) from Escherichia coli was determined via the multiwavelength anomalous diffraction technique using a selenomethionine-labeled crystal containing 88 selenium sites in the asymmetric unit. The comparison of the E. coli FabG structure with the homologous Brassica napus FabG.NADP(+) binary complex reveals that cofactor binding causes a substantial conformational change in the protein. This conformational change puts all three active-site residues (Ser 138, Tyr 151, and Lys 155) into their active configurations and provides a structural mechanism for allosteric communication between the active sites in the homotetramer. FabG exhibits negative cooperative binding of NADPH, and this effect is enhanced by the presence of acyl carrier protein (ACP). NADPH binding also increases the affinity and decreases the maximum binding of ACP to FabG. Thus, unlike other members of the short-chain dehydrogenase/reductase superfamily, FabG undergoes a substantial conformational change upon cofactor binding that organizes the active-site triad and alters the affinity of the other substrate-binding sites in the tetrameric enzyme.  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