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 15 毫秒
1.
The flavoprotein Escherichia coli methylenetetrahydrofolate reductase (MTHFR) catalyzes the reduction of 5,10-methylenetetrahydrofolate (CH(2)-H(4)folate) to 5-methyltetrahydrofolate (CH(3)-H(4)folate). The X-ray crystal structure of the enzyme has revealed the amino acids at the flavin active site that are likely to be relevant to catalysis. Here, we have focused on two conserved residues, Asp 120 and Glu 28. The presence of an acidic residue (Asp 120) near the N1-C2=O position of the flavin distinguishes MTHFR from all other known flavin oxidoreductases and suggests an important function for this residue in modulating the flavin reactivity. Modeling of the CH(3)-H(4)folate product into the enzyme active site also suggests roles for Asp 120 in binding of folate and in electrostatic stabilization of the putative 5-iminium cation intermediate during catalysis. In the NADH-menadione oxidoreductase assay and in the isolated reductive half-reaction, the Asp120Asn mutant enzyme is reduced by NADH 30% more rapidly than the wild-type enzyme, which is consistent with a measured increase in the flavin midpoint potential. Compared to the wild-type enzyme, the mutant showed 150-fold decreased activity in the physiological NADH-CH(2)-H(4)folate oxidoreductase reaction and in the oxidative half-reaction involving CH(2)-H(4)folate, but the apparent K(d) for CH(2)-H(4)folate was relatively unchanged. Our results support a role for Asp 120 in catalysis of folate reduction and perhaps in stabilization of the 5-iminium cation. By analogy to thymidylate synthase, which also uses CH(2)-H(4)folate as a substrate, Glu 28 may serve directly or via water as a general acid catalyst to aid in 5-iminium cation formation. Consistent with this role, the Glu28Gln mutant was unable to catalyze the reduction of CH(2)-H(4)folate and was inactive in the physiological oxidoreductase reaction. The mutant enzyme was able to bind CH(3)-H(4)folate, but reduction of the FAD cofactor was not observed. In the NADH-menadione oxidoreductase assay, the mutant demonstrated a 240-fold decrease in activity.  相似文献   

2.
Elevated plasma homocysteine levels are associated with increased risk for cardiovascular disease and neural tube defects in humans. Folate treatment decreases homocysteine levels and dramatically reduces the incidence of neural tube defects. The flavoprotein methylenetetrahydrofolate reductase (MTHFR) is a likely target for these actions of folate. The most common genetic cause of mildly elevated plasma homocysteine in humans is the MTHFR polymorphism A222V (base change C677-->T). The X-ray analysis of E. coli MTHFR, reported here, provides a model for the catalytic domain that is shared by all MTHFRs. This domain is a beta8alpha8 barrel that binds FAD in a novel fashion. Ala 177, corresponding to Ala 222 in human MTHFR, is near the bottom of the barrel and distant from the FAD. The mutation A177V does not affect Km or k(cat) but instead increases the propensity for bacterial MTHFR to lose its essential flavin cofactor. Folate derivatives protect wild-type and mutant E. coli enzymes against flavin loss, and protect human MTHFR and the A222V mutant against thermal inactivation, suggesting a mechanism by which folate treatment reduces homocysteine levels.  相似文献   

3.
4.
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   

5.
Sequence-specific 1H assignments have been made for over 25% of the amino acid side chains of Escherichia coli dihydrofolate reductase complexed with folate by using a variety of two-dimensional techniques. Proton resonances were assigned by using a combination of site-directed mutagenesis and a knowledge of the X-ray crystal structure. Unique sets of NOE connectivities present in hydrophobic pockets were matched with the X-ray structure and used to assign many of the residues. Other residues, particularly those near or in the active site, were assigned by site-directed mutagenesis. The ability to assign unambiguously the proton resonances of these catalytically important residues allowed for extensive networks of NOE connectivities to follow from these assignments. As a consequence of these assignments, the orientation of the pterin ring of folate could be determined, and its conformation is similar to that of the productive dihydrofolate complex. Under these experimental conditions, only one bound form of the pterin ring could be detected.  相似文献   

6.
S Zenno  H Koike  M Tanokura    K Saigo 《Journal of bacteriology》1996,178(15):4731-4733
NfsB is an oxygen-insensitive nitroreductase of Escherichia coli with significant amino acid sequence homology to the major flavin reductase (FRase I) of Vibrio fischeri. Here, we show that NfsB is convertible to an FRase I-like flavin reductase three times as active as the authentic FRase I by a single amino acid substitution in the least-conserved region.  相似文献   

7.
Pejchal R  Sargeant R  Ludwig ML 《Biochemistry》2005,44(34):11447-11457
Methylenetetrahydrofolate reductases (MTHFRs; EC 1.7.99.5) catalyze the NAD(P)H-dependent reduction of 5,10-methylenetetrahydrofolate (CH(2)-H(4)folate) to 5-methyltetrahydrofolate (CH(3)-H(4)folate) using flavin adenine dinucleotide (FAD) as a cofactor. The initial X-ray structure of Escherichia coli MTHFR revealed that this 33-kDa polypeptide is a (betaalpha)(8) barrel that aggregates to form an unusual tetramer with only 2-fold symmetry. Structures of reduced enzyme complexed with NADH and of oxidized Glu28Gln enzyme complexed with CH(3)-H(4)folate have now been determined at resolutions of 1.95 and 1.85 A, respectively. The NADH complex reveals a rare mode of dinucleotide binding; NADH adopts a hairpin conformation and is sandwiched between a conserved phenylalanine, Phe223, and the isoalloxazine ring of FAD. The nicotinamide of the bound pyridine nucleotide is stacked against the si face of the flavin ring with C4 adjoining the N5 of FAD, implying that this structure models a complex that is competent for hydride transfer. In the complex with CH(3)-H(4)folate, the pterin ring is also stacked against FAD in an orientation that is favorable for hydride transfer. Thus, the binding sites for the two substrates overlap, as expected for many enzymes that catalyze ping-pong reactions, and several invariant residues interact with both folate and pyridine nucleotide substrates. Comparisons of liganded and substrate-free structures reveal multiple conformations for the loops beta2-alpha2 (L2), beta3-alpha3 (L3), and beta4-alpha4 (L4) and suggest that motions of these loops facilitate the ping-pong reaction. In particular, the L4 loop adopts a "closed" conformation that allows Asp120 to hydrogen bond to the pterin ring in the folate complex but must move to an "open" conformation to allow NADH to bind.  相似文献   

8.
The importance of salt bridge interactions at the NADPH binding site of dihydrofolate reductase has been studied by using site-directed mutagenesis. The mutations R44L and H45Q respectively disrupt the ionic contacts made between the 2'-phosphate and pyrophosphoryl moiety of the coenzyme and the N-terminal region of helix C. Equilibrium fluorescence experiments indicate that while the overall binding of NADPH to both free mutants is weakened by 1.1 and 1.5 kcal/mol (H45Q and R44L, respectively), the binding of dihydrofolate and tetrahydrofolate is unaffected. Despite the similar binding energies for both mutants, the transition state for the chemical hydride step is differentially destabilized relative to wild type (0.6 and 1.8 kcal/mol for H45Q and R44L, respectively). Both stopped-flow and pre-steady-state experiments suggest that the root of this effect may lie in multiple conformations for the E-NADPH complex of R44L. The ability of both mutants to transmit their effects beyond the local environment of the NADPH pocket is manifested in several details: (1) the pKa of Asp-27 (25 A away from the sites of mutation) is elevated from 6.5 in the wild type to 7.5 and 8.4 in H45Q and R44L, respectively; (2) NADPH elevates the off rates for tetrahydrofolate from 12 s-1 in the wild type to greater than 45 s-1 in R44L; and (3) bound tetrahydrofolate decreases the affinity of the enzymes for NADPH as reflected in the Km from 2 to 40 microM for H45Q (similar to wild type) but from 8 to 5000 microM for R44L.  相似文献   

9.
The role of Thr-113 of Escherichia coli dihydrofolate reductase in binding and catalysis was probed by amino acid substitution. Thr-113, a strictly conserved residue that forms a hydrogen bond to the active-site Asp-27 and to the amino group of methotrexate through a fixed water molecule, was replaced by valine. The kinetic scheme is identical in form with the wild-type scheme, although many of the rate constants vary, including a decrease in the association rate constants and an increase in the dissociation rate constants for folate ligands, a decrease in the hydride-transfer rate constant in both directions, and an increase in the intrinsic pKa of Asp-27. Overall, replacement of Thr-113 by Val decreases the binding of folate substrates by approximately 2.3 kcal/mol. These multiple complex changes on various ground and transition states underscore the optimal properties of a strictly conserved residue in the evolution of catalytic function.  相似文献   

10.
L C Kuo  A W Miller  S Lee  C Kozuma 《Biochemistry》1988,27(24):8823-8832
In the carbamoyl-transfer reaction catalyzed by ornithine transcarbamoylase, an arginine residue in the active site of the Escherichia coli enzyme has been suggested to bind the phosphate moiety of the substrate carbamoyl phosphate. With the application of site-specific mutagenesis, the most likely arginine residue among three candidates at the binding site of carbamoyl phosphate, Arg-57, has been replaced with a glycine. The resultant Gly-57 mutant enzyme is drastically inefficient in catalysis. In the synthesis of L-citrulline from carbamoyl phosphate and L-ornithine with the release of inorganic phosphate, the turnover rate of the mutant is 21,000-fold lower than that of the wild type. However, the mutation of Arg-57 affects only moderately the binding of carbamoyl phosphate; the dissociation constant of this substrate, measured under steady-state turnover condition, is increased from 0.046 to 3.2 mM by the mutation. On the other hand, ornithine binding is substantially affected as estimated by the change in the dissociation constant of its analogue L-norvaline. The dissociation constant of L-norvaline increases about 500-fold from 54 microM for the wild type to 25 mM for the mutant. Since Arg-57 is expected to be distal from the ornithine site and the amino acid (both ornithine and norvaline) binds only after carbamoyl phosphate in the wild-type reaction, the poor norvaline affinity to the mutant suggests that Arg-57 is involved in interactions essential for productive addition of the amino acid. This interpretation is supported by difference ultraviolet absorption spectra which show that the conformational changes induced in the wild type by carbamoyl phosphate upon binding are absent in the mutant. Furthermore, steady-state kinetic data reveal that the ordered binding mechanism of the wild-type enzyme is transformed into a random binding mechanism in the mutant. Thus, the presence of carbamoyl phosphate in the mutant active site is no longer a requisite for ornithine binding. In the 5-50 degrees C temperature range, transcarbamoylation catalyzed by either the wild type or the mutant observes the Arrhenius rate law with almost identical enthalpies of activation, 11 and 10 kcal/mol, respectively. The entropy of activation is -5.5 eu for the wild-type reaction and -29 eu for the mutant reaction, accounting for a loss of 6-7 kcal/mol in the rate-determining step of the enzymic reaction.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
Membrane-bound nitrate reductase of Escherichia coli consists of three subunits designated as A, B, and C, with subunit C being the apoprotein of cytochrome b, A hemA mutant that cannot synthesize delta-aminolevulinic acid (ALA) produces a normal, stable, membrane-bound enzyme when grown with ALA. When grown without ALA, this mutant makes a reduced amount of membrane-bound enzyme that is unstable and contains no C subunit. Under the same growth conditions, this mutant accumulates a large amount of a soluble form of the enzyme in the cytoplasm. Accumulation of this cytoplasmic form begins immediately upon induction of the enzyme with nitrate. The cytoplasmic form is very similar to the soluble form of the enzyme obtained by alkaline heat extraction. It is a high-molecular-weight complex with a Strokes radius of 8.0 nm and consists of intact A and B subunits. When ALA is added to a culture growing without ALA, the cytoplasmic form of the enzyme is incorporated into the membrane in a stable form, coincident with the formation of functional cytochrome b. Reconstitution experiments indicate that subunit C is present in cultures grown without ALA but is reduced in amount or unstable. These results indicate that membrane-bound nitrate reductase is synthesized via a soluble precursor containing subunits A and B, which then binds to the membrane upon interaction with the third subunit, cytochrome b.  相似文献   

12.
Subunits A and B were isolated from purified nitrate reductase by preparative electrophoresis in low levels of sodium dodecyl sulfate. Nonheme iron and low levels of molybdenum were associated with isolated subunit A but not with isolated subunit B. After dialysis against a source of molybdenum cofactor, subunit A regained tightly bound molybdenum and concomitantly regained enzyme activity and reactivity with anti-nitrate reductase antiserum. Subunit B neither bound cofactor nor regained activity or reactivity with antiserum. These data indicate that subunit A contains the active site of the enzyme. Subunit A was also found to be modified posttranslationally in a similar fashion as is subunit B. This was determined by comparison of partial proteolytic digests and amino acid analyses of A subunits from precursor and membrane-bound forms of nitrate reductase.  相似文献   

13.
14.
15.
The Escherichia coli complex II homologues succinate:ubiquinone oxidoreductase (SQR, SdhCDAB) and menaquinol:fumarate oxidoreductase (QFR, FrdABCD) have remarkable structural homology at their dicarboxylate binding sites. Although both SQR and QFR can catalyze the interconversion of fumarate and succinate, QFR is a much better fumarate reductase, and SQR is a better succinate oxidase. An exception to the conservation of amino acids near the dicarboxylate binding sites of the two enzymes is that there is a Glu (FrdA Glu-49) near the covalently bound FAD cofactor in most QFRs, which is replaced with a Gln (SdhA Gln-50) in SQRs. The role of the amino acid side chain in enzymes with Glu/Gln/Ala substitutions at FrdA Glu-49 and SdhA Gln-50 has been investigated in this study. The data demonstrate that the mutant enzymes with Ala substitutions in either QFR or SQR remain functionally similar to their wild type counterparts. There were, however, dramatic changes in the catalytic properties when Glu and Gln were exchanged for each other in QFR and SQR. The data show that QFR and SQR enzymes are more efficient succinate oxidases when Gln is in the target position and a better fumarate reductase when Glu is present. Overall, structural and catalytic analyses of the FrdA E49Q and SdhA Q50E mutants suggest that coulombic effects and the electronic state of the FAD are critical in dictating the preferred directionality of the succinate/fumarate interconversions catalyzed by the complex II superfamily.  相似文献   

16.
Complex II superfamily members catalyze the kinetically difficult interconversion of succinate and fumarate. Due to the relative simplicity of complex II substrates and their similarity to other biologically abundant small molecules, substrate specificity presents a challenge in this system. In order to identify determinants for on-pathway catalysis, off-pathway catalysis, and enzyme inhibition, crystal structures of Escherichia coli menaquinol:fumarate reductase (QFR), a complex II superfamily member, were determined bound to the substrate, fumarate, and the inhibitors oxaloacetate, glutarate, and 3-nitropropionate. Optical difference spectroscopy and computational modeling support a model where QFR twists the dicarboxylate, activating it for catalysis. Orientation of the C2-C3 double bond of activated fumarate parallel to the C(4a)-N5 bond of FAD allows orbital overlap between the substrate and the cofactor, priming the substrate for nucleophilic attack. Off-pathway catalysis, such as the conversion of malate to oxaloacetate or the activation of the toxin 3-nitropropionate may occur when inhibitors bind with a similarly activated bond in the same position. Conversely, inhibitors that do not orient an activatable bond in this manner, such as glutarate and citrate, are excluded from catalysis and act as inhibitors of substrate binding. These results support a model where electronic interactions via geometric constraint and orbital steering underlie catalysis by QFR.  相似文献   

17.
18.
Studies reported here were designed to examine the hypothesis that covalent binding of reactive intermediates to macromolecules of the conceptus represents a major mechanism for the embryotoxicity of niridazole (NDZ). The roles of embryonic thiol content and oxygenation on: 1) malformation incidence; 2) reductive metabolism; and 3) covalent binding to embryonic macromolecules of metabolites resulting from reductive biotransformation of NDZ were studied. Results were compared with those from studies with the nondysmorphogenic analog of NDZ, 4'-methylniridazole (MNDZ). Day 10 rat embryos were pretreated for 5 hours in vitro with either L-buthionine-S, R-sulfoximine (BSO) or N-acetylcysteine (NAC) to modulate their glutathione (GSH) content. BSO reduced GSH levels, but NAC was ineffective. Following pretreatment, embryos were cultured for an additional 15 hours in the presence of [14C]NDZ or [14C]MNDZ with an initial oxygen concentration of 5%. At the end of the culture period (day 11, AM), those embryos with active heartbeat and vitelline circulation were examined for asymmetric malformations. Drug metabolites were subjected to multiple extractions from the culture medium and subjected to quantitative high-performance liquid chromatography (HPLC) analysis. Homogenates of the embryos were extracted with trichloroacetic acid (TCA) to estimate the covalent binding of radiolabeled parent compound/metabolites. Autoradiographic analyses were performed on other embryos. BSO pretreatment, which reduces embryonic GSH tissue levels, dramatically increased both the conversion of NDZ to 1-thiocarbamoyl-2-imidazolidinone (TCI) (generated via reductive metabolism of NDZ) and covalently bound label but failed to increase embryotoxicity. NAC, by contrast, did not significantly affect embryonic GSH levels, TCI generation, or covalent binding. Because both rates of metabolism of NDZ to TCI and covalent binding could vary independently of malformation incidence, we concluded that they do not represent critical mechanistic factors for the embryotoxicity of NDZ and related nitroheterocycles.  相似文献   

19.
5-DeazaFAD bound to a hydrophobic site in apophotolyase and formed a stable reconstituted enzyme, similar to that observed with FAD. Although stoichiometric incorporation was observed, the flavin ring modification in 1-deazaFAD interfered with normal binding, decreased protein stability, and prevented formation of a stable flavin radical, unlike that observed with FAD. The results suggest that an important hydrogen bond is formed between the protein and N (1) in FAD, but not N (5), and that there is sufficient space at the normal flavin binding site near N (5) to accommodate an additional hydrogen but not near N (1). Catalytic activity was observed with enzyme containing 5-deazaFADH2 (42% of native enzyme) or 1-deazaFADH2 (11% of native enzyme) as its only chromophore, but no activity was observed with the corresponding oxidized flavins, similar to that observed with FAD and consistent with a mechanism where dimer cleavage is initiated by electron donation from excited reduced flavin to substrate. The protein environment in photolyase selectively enhanced photochemical reactivity in the fully reduced state, as evidenced by comparison with results obtained in model studies with the corresponding free flavins. Phosphorescence was observed with free or photolyase-bound 5-deazaFADH2, providing the first example of a flavin that exhibits phosphorescence in the fully reduced state. Formation of an enzyme-substrate complex resulted in a nearly identical extent of quenching of 5-deazaFADH2 phosphorescence (85.1%) and fluorescence (87.5%). The data are consistent with a mechanism involving exclusive reaction of substrate with the excited singlet state of 5-deazaFADH2, analogous to that proposed for FADH2 in native enzyme. Direct evidence for singlet-singlet energy transfer from enzyme-bound 5-deazaFADH2 to 5,10-CH(+)-H4folate was provided by the fact that pterin fluorescence was observed upon excitation of 5-deazaFADH2, accompanied by a decrease in 5-deazaFADH2 fluorescence. On the other hand, the fluorescence of enzyme-bound pterin was quenched by 5-deazaFADox, consistent with energy transfer from pterin to 5-deazaFADox. In each case, the spectral properties of the chromophores were consistent with the observed direction of energy transfer and indicated that transfer in the opposite direction was energetically unlikely. Unlike 5-deazaFAD, energy transfer from pterin to FAD is energetically feasible with FADH2 or FADox. The results indicate that the direction of flavin-pterin energy transfer at the active site of photolyase can be manipulated by changes in the flavin ring or redox state which alter the energy level of the flavin singlet.  相似文献   

20.
C Bystroff  S J Oatley  J Kraut 《Biochemistry》1990,29(13):3263-3277
The crystal structure of dihydrofolate reductase (EC 1.5.1.3) from Escherichia coli has been solved as the binary complex with NADP+ (the holoenzyme) and as the ternary complex with NADP+ and folate. The Bragg law resolutions of the structures are 2.4 and 2.5 A, respectively. The new crystal forms are nonisomorphous with each other and with the methotrexate binary complex reported earlier [Bolin, J. T., Filman, D. J., Matthews, D. A., Hamlin, R. C., & Kraut, J. (1982) J. Biol. Chem. 257, 13650-13662]. In general, NADP+ and folate binding conform to predictions, but the nicotinamide moiety of NADP+ is disordered in the holoenzyme and ordered in the ternary complex. A mobile loop (residues 16-20) involved in binding the nicotinamide is also disordered in the holoenzyme. We report a detailed analysis of the binding interactions for both ligands, paying special attention to several apparently strained interactions that may favor the transition state for hydride transfer. Hypothetical models are presented for the binding of 7,8-dihydrofolate in the Michaelis complex and for the transition-state complex.  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