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 flavin-containing enzyme dihydroorotate dehydrogenase (DHOD) catalyzes the oxidation of dihydroorotate (DHO) to orotate, the first aromatic intermediate in pyrimidine biosynthesis. The first structure of a DHOD, the A form of the enzyme from Lactococcus lactis, has recently become known, and some conserved residues were suggested to have a role in the active site [Rowland et al. (1997) Structure 2, 239-252]. In particular, Cys 130 was hypothesized to work as a base, which activates dihydroorotate (DHO) for hydride transfer. By chemical modification and site-directed mutagenesis we have obtained results consistent with this proposal. Cys 130 was susceptible to alkylating reagents, and mutants of Cys 130 (C130A and C130S) showed hardly detectable enzyme activity at pH 8.0, while at pH 10 the C130S mutant enzyme had approximately 1% of wild-type activity. Mutants of Lys 43, Asn 132, and Lys 164 were also constructed. Exchange of Lys 43 to Ala or Glu (K43A and K43E) and of Asn 132 to Ala (N132A) affected both catalysis and substrate binding. Expressed as kcat/KM for DHO, the deterioration of these three mutant enzymes was 10(3)-10(4)-fold. Flavin spectra of the mutant enzymes were not, like the wild-type enzyme, bleached by DHO in stopped-flow experiments, showing that they were deficient with respect to the first half-reaction, namely reduction of FMN by DHO, which was not rate limiting for the wild-type enzyme. The binding interaction between flavin and the reaction product, orotate, could be monitored by a red shift of the flavin absorbance in the wild-type enzyme. The C130A, C130S, and N132A mutant enzymes displayed similar capacity to bind orotate. In contrast, orotate did not change the absorption spectra of the K43 mutant enzymes, although it did inhibit their activity. All of the mutant enzymes, except K164A, contained normal levels of flavin. The results are discussed in relation to the structures of DHODA and other flavoenzymes. The possible acid-base chemistry of Cys 130 is compared to previous work on mammalian dihydropyrimidine dehydrogenases, flavoenzymes, which catalyze the reversed reaction, namely the reduction of pyrimidine bases.  相似文献   

2.
Nitric oxide synthase (NOS) catalyzes the oxidation of L-arginine to citrulline and nitric oxide. C415H and C415A mutants of the neuronal isoform of NOS (nNOS) were expressed in a baculovirus system and purified to homogeneity for spectral analysis and activity measurements. UV-visible spectra of each mutant lacked an observable Soret peak, suggesting that neither mutant contained heme. When reduced in the presence of CO, however, a small Soret centered at 417 nm could be detected for the C415H mutant, further supporting the assignment of C415 as the axial ligand to the heme. In addition to a deficiency in bound heme, neither mutant had any detectable bound tetrahydrobiopterin, as compared to wild-type enzyme, which had a ratio of 0.84 mol of bound pteridine:1 mol of nNOS 160 kDa subunit. The C415H mutant contained bound FAD and FMN at levels of 1.0 +/- 0.1 and 0.9 +/- 0.1 mol/mol of nNOS subunit, respectively. UV-visible spectra of both nNOS mutants retained the distinctive absorbance due to tightly associated oxidized flavin prosthetic groups. Further, the spectra suggested the presence of a neutral flavin semiquinone. Ferricyanide oxidation of the C415A mutant yielded a spectrum that was essentially that of oxidized flavin. Ferricyanide titration showed that the C415A mutant contained approximately 1 reducing equiv. Circular dichroism spectra suggested that each mutant was folded properly, in that both spectra were found to be essentially identical to the spectrum of wild-type nNOS. Neither mutant could synthesize nitric oxide, and neither mutant had the ability to oxidize NADPH unless an exogenous electron acceptor was added. The rate of cytochrome c reduction by each mutant was found to be slightly less, but very similar to the rate (approximately 20 mumol mg-1 min-1) observed with wild-type nNOS. In all cases, the rate of cytochrome c reduction increased approximately 15-fold with the addition of calmodulin. Overall, these spectral and activity data suggest that C415 is the axial heme ligand and that a point mutation at C415 prevents binding of heme and tetrahydrobiopterin without interfering with the global folding or the reductase function of nNOS.  相似文献   

3.
Old Yellow Enzyme (OYE) binds phenolic ligands forming long wavelength (500-800 nm) charge-transfer complexes. The enzyme is reduced by NADPH, and oxygen, quinones, and alpha,beta-unsaturated aldehydes and ketones can act as electron acceptors to complete catalytic turnover. Solution of the crystal structure of OYE1 from brewer's bottom yeast (Fox, K. M., and Karplus, P. A. (1994) Structure 2, 1089-1105) made it possible to identify histidine 191 and asparagine 194 as amino acid residues that hydrogen-bond with the phenolic ligands, stabilizing the anionic form involved in charge-transfer interaction with the FMN prosthetic group. His-191 and Asn-194 are also predicted to interact with the nicotinamide ring of NADPH in the active site. Mutations of His-191 to Asn, Asn-194 to His, and a double mutation, H191N/N194H, were made of OYE1. It was not possible to isolate the N191H mutant enzyme, but the other two mutant forms had the expected effect on phenolic ligand binding, i.e. decreased binding affinity and decreased charge-transfer absorbance. Reduction of the H191N mutant enzyme by NADPH was similar to that of OYE1, but the reduction rate constant for NADH was greatly decreased. The double mutant enzyme had an increased rate constant for reduction by NADPH, but the reduction rate constant with NADH was lower by a factor of 15. The reactivity of OYE1 and the mutant enzymes with oxygen was similar, but the reactivity of 2-cyclohexenone was greatly decreased by the mutations. The crystal structures of the two mutant forms showed only minor changes from that of the wild type enzyme.  相似文献   

4.
Deficiency of dihydropyrimidine dehydrogenase or dihydropyrimidinase, enzymes that catalyze the breakdown of pyrimidine chemotherapy agents such as 5-fluorouracil, may cause serious adverse reactions to these agents. We attempted to establish the reference range for urinary pyrimidines in adults to detect individuals with abnormal pyrimidine metabolism. We analyzed urinary pyrimidine levels in 1133 adults to establish a reference range for persons ages 20 years or older. Urinary dihydrouracil and uracil levels were determined by high-performance liquid chromatography with column switching. The reference range obtained was found to be 0-59.3 micromol/g creatinine for dihydrouracil and 0-129.8 micromol/g creatinine for uracil. In addition, an asymptomatic man with suspected dihydropyrimidinase deficiency was detected on the basis of dihydropyrimidinuria. Although only three cases of this disease have been found worldwide, including one infant reported previously by our group, it may not be so rare as has been thought. In this man, a 10 mg/kg oral uracil loading test yielded a peak blood dihydrouracil level of 192.1 micromol/liter and a peak uracil level of 67.8 micromol/liter. Eight h after loading, the uracil level was still 11.1 micromol/liter, about 17 times that in healthy subjects. Additional research on dihydropyrimininase deficiency may help to prevent adverse reactions to pyrimidine chemotherapy agents in susceptible individuals.  相似文献   

5.
The human pathogen Staphylococcus aureus does not utilize the glutathione thiol/disulfide redox system employed by eukaryotes and many bacteria. Instead, this organism produces CoA as its major low molecular weight thiol. We report the identification and purification of the disulfide reductase component of this thiol/disulfide redox system. Coenzyme A disulfide reductase (CoADR) catalyzes the specific reduction of CoA disulfide by NADPH. CoADR has a pH optimum of 7.5-8.0 and is a dimer of identical subunits of Mr 49,000 each. The visible absorbance spectrum is indicative of a flavoprotein with a lambdamax = 452 nm. The liberated flavin from thermally denatured enzyme was identified as flavin adenine dinucleotide. Steady-state kinetic analysis revealed that CoADR catalyzes the reduction of CoA disulfide by NADPH at pH 7.8 with a Km for NADPH of 2 muM and for CoA disulfide of 11 muM. In addition to CoA disulfide CoADR reduces 4,4'-diphosphopantethine but has no measurable ability to reduce oxidized glutathione, cystine, pantethine, or H2O2. CoADR demonstrates a sequential kinetic mechanism and employs a single active site cysteine residue that forms a stable mixed disulfide with CoA during catalysis. These data suggest that S. aureus employs a thiol/disulfide redox system based on CoA/CoA-disulfide and CoADR, an unorthodox new member of the pyridine nucleotide-disulfide reductase superfamily.  相似文献   

6.
To investigate the functional role of the cysteine residues present in the spinach ferredoxin-NADP+ oxidoreductase, we individually replaced each of the five cysteine residues with serine using site-directed mutagenesis. All of the mutant reductases were correctly assembled in Escherichia coli except for the C42S mutant protein. C114S and C137S mutant enzymes apparently showed structural and kinetic properties very similar to those of the wild-type reductase. However, C272S and C132S mutations yielded enzymes with a decreased catalytic activity in the ferredoxin-dependent reaction (14 and 31% of the wild type, respectively). Whereas the C132S was fully competent in the diaphorase reaction, the C272S mutant flavoprotein showed a 35-fold reduction in catalytic efficiency with respect to the wild-type enzyme (0.4 versus 14.28 microM-1 s-1) due to a substantial decrease of kcat. NADP+ binding by the C272S mutant enzyme was apparently quantitatively the same (Kd = 37 microM) but qualitatively different, as shown by the differential spectrum. Stopped-flow experiments showed that the enzyme-FAD reduction rate was considerably decreased in the C272S mutant reductase, along with a much lower yield of the charge-transfer transient species. It is inferred from these data that the charge transfer (FAD-NADPH) between the reductase and NADPH is required for hydride transfer from the pyridine nucleotide to flavin to occur with a rate compatible with catalysis.  相似文献   

7.
The isoalloxazine ring system of the FAD cofactor of p-hydroxybenzoate hydroxylase must be secluded from solvent at specific stages of catalysis in order to form and stabilize a flavin C4a-hydroperoxide. This species may then react with the activated phenolate of p-hydroxybenzoate. A number of crystal structures of the enzyme with alterations to active site substituents or complexes with analogue benzoates have revealed an alternate position for the isoalloxazine (Gatti et al. (1994) Science 266, 110-114; Schreuder et al. (1994) Biochemistry 33, 10161-10170). This new flavin conformation is 7 A "out" toward solvent and may open a passage for substrate entry to the active site. Arginine 220 is one of the few residues in the structure to demonstrate conformational changes when the flavin is "out". In this study we have made the Arg220Lys mutant to test the significance of this residue in flavin movement. The R220K mutation has brought about dramatic alterations to all aspects of catalysis. Stopped flow kinetic characterization of the mutant has revealed that, while the effector role for the substrate is maintained, there exists an order of magnitude decrease in the limiting rate of reduction, even though there is 40-fold increase in association with NADPH. The mutant enzyme has only a fraction of its reductive half-reaction coupled to product formation, and the hydroxylation process is slow. This occurs despite a higher proportion of the more activated substrate phenolate in the active site. Many of the observed changes can be attributed to a decrease in the stability of the "in" conformation of the flavin during the catalysis and indicate a role for flavin conformational states in many of the catalytic processes of the enzyme.  相似文献   

8.
Crystallographic studies have demonstrated two flavin conformations for p-hydroxybenzoate hydroxylase (PHBH) [Gatti, D. L., Palfey, B. A. , Lah, M. S., Entsch, B., Massey, V., Ballou, D. P., & Ludwig, M. L. (1994) Science 266, 110-114. Schreuder, H. A., Mattevi, A., Obmolova, G., Kalk, K. H., Hol, W. G. J., van der Bolt, F. J. T., & van Berkel, W. J. H. (1994) Biochemistry 33, 10161-10170]. The isoalloxazine ring system of one conformation (the "out" conformation) is significantly more exposed to solvent and is not in position for necessary catalytic reactions, but when the natural substrate is bound to the enzyme, the isoalloxazine is in the correct position (the "in" conformation) for its chemical function. In this study, several aspects of the function of the conformational change in catalysis were explored using the wild-type and Tyr222Phe forms of PHBH substituted with 6-azido FAD. This flavin served as both a spectral probe and a photolabel. The enzyme containing 6-azido FAD was a relatively effective catalyst for the hydroxylation of p-hydroxybenzoate. However, the intermediate reduced 6-azido enzyme was chemically unstable, and a small fraction converted to 6-amino PHBH by the elimination of N2 during each catalytic cycle. The reduction of 6-azido FAD PHBH by NADPH was almost as fast as the reduction of the natural enzyme. The characteristic spectral change caused by NADPH binding prior to hydride transfer strongly suggests that flavin movement from the "in" to the "out" conformation precedes flavin reduction. Irradiation of 6-azido PHBH with visible light covalently labeled proline 293, an active site residue, under conditions in which the flavin adopted the "in" conformation, while no protein labeling occurred under conditions in which the flavin was "out". The labeled protein exchanged substrate and was reduced by NADPH much more slowly than before photolysis. It is therefore concluded that isoalloxazine movement is required for pyridine nucleotide to gain access to the active site and for the exchange of aromatic ligands.  相似文献   

9.
The enzymatic properties and the three-dimensional structure of spinach glycolate oxidase which has the active-site Tyr129 replaced by Phe (Y129F glycolate oxidase) has been studied. The structure of the mutant is unperturbed which facilitates interpretation of the biochemical data. Y129F glycolate oxidase has an absorbance spectrum with maxima at 364 and 450 nm (epsilon max = 11400 M-1 cm-1). The spectrum indicates that the flavin is in its normal protonated form, i.e. the Y129F mutant does not lower the pKa of the N(3) of oxidized flavin as does the wild-type enzyme [Macheroux, P., Massey, V., Thiele, D. J., and Volokita, M. (1991) Biochemistry 30, 4612-4619]. This was confirmed by a pH titration of Y129F glycolate oxidase which showed that the pKa is above pH 9. In contrast to wild-type glycolate oxidase, oxalate does not perturb the absorbance spectrum of Y129F glycolate oxidase. Moreover oxalate does not inhibit the enzymatic activity of the mutant enzyme. Typical features of wild-type glycolate oxidase that are related to a positively charged lysine side chain near the flavin N(1)-C(2 = O), such as stabilization of the anionic flavin semiquinone and formation of tight N(5)-sulfite adducts, are all conserved in the Y129F mutant protein. Y129F glycolate oxidase exhibited about 3.5% of the wild-type activity. The lower turnover number for the mutant of 0.74 s-1 versus 20 s-1 for the wild-type enzyme amounts to an increase of the energy of the transition state of about 7.8 kJ/mol. Steady-state analysis gave Km values of 1.5 mM and 7 microM for glycolate and oxygen, respectively. The Km for glycolate is slightly higher than that found for wild-type glycolate oxidase (1 mM) whereas the Km for oxygen is much lower. As was the case for wild-type glycolate oxidase, reduction was found to be the rate-limiting step in catalysis, with a rate of 0.63 s-1. The kinetic properties of Y129F glycolate oxidase provide evidence that the main function of the hydroxyl group of Tyr129 is the stabilization of the transition state.  相似文献   

10.
Comparisons of the structures of glutamate dehydrogenase (GluDH) and leucine dehydrogenase (LeuDH) have suggested that two substitutions, deep within the amino acid binding pockets of these homologous enzymes, from hydrophilic residues to hydrophobic ones are critical components of their differential substrate specificity. When one of these residues, K89, which hydrogen-bonds to the gamma-carboxyl group of the substrate l-glutamate in GluDH, was altered by site-directed mutagenesis to a leucine residue, the mutant enzyme showed increased substrate activity for methionine and norleucine but negligible activity with either glutamate or leucine. In order to understand the molecular basis of this shift in specificity we have determined the crystal structure of the K89L mutant of GluDH from Clostridium symbiosum. Analysis of the structure suggests that further subtle differences in the binding pocket prevent the mutant from using a branched hydrophobic substrate but permit the straight-chain amino acids to be used as substrates. The three-dimensional crystal structure of the GluDH from C. symbiosum has been previously determined in two distinct forms in the presence and absence of its substrate glutamate. A comparison of these two structures has revealed that the enzyme can adopt different conformations by flexing about the cleft between its two domains, providing a motion which is critical for orienting the partners involved in the hydride transfer reaction. It has previously been proposed that this conformational change is triggered by substrate binding. However, analysis of the K89L mutant shows that it adopts an almost identical conformation with that of the wild-type enzyme in the presence of substrate. Comparison of the mutant structure with both the wild-type open and closed forms has enabled us to separate conformational changes associated with substrate binding and domain motion and suggests that the domain closure may well be a property of the wild-type enzyme even in the absence of substrate.  相似文献   

11.
Tyr114 and Tyr197 are highly conserved residues in the active site of human glutathione reductase, Tyr114 in the glutathione disulfide (GSSG) binding site and Tyr197 in the NADPH site. Mutation of either residue has profound effects on catalysis. Y197S and Y114L have 17% and 14% the activity of the wild-type enzyme, respectively. Mutation of Tyr197, in the NADPH site, leads to a decrease in Km for GSSG, and mutation of Tyr114, in the GSSG site, leads to a decrease in Km for NADPH. This behavior is predicted for enzymes operating by a ping-pong mechanism where both half-reactions partially limit turnover. Titration of the wild-type enzyme or Y114L with NADPH proceeds in two phases, Eox to EH2 and EH2 to EH2-NADPH. In contrast, Y197S reacts monophasically, showing that excess NADPH fails to enhance the absorbance of the thiolate-FAD charge-transfer complex, the predominant EH2 form of glutathione reductase. The reductive half-reactions of the wild-type enzyme and of Y114L are similar; FAD reduction is fast (approximately 500 s-1 at 4 degreesC) and thiolate-FAD charge-transfer complex formation has a rate of 100 s-1. In Y197S, these rates are only 78 and 5 s-1, respectively. The oxidative half-reaction, the rate of reoxidation of EH2 by GSSG, of the wild-type enzyme is approximately 4-fold faster than that of Y114L. These results are consistent with Tyr197 serving as a gate in the binding of NADPH, and they indicate that Tyr114 assists the acid catalyst His467'.  相似文献   

12.
L-Lactate dehydrogenase from Bifidobacterium longum shows homotropic activation by pyruvate as well as heterotropic activation by fructose 1,6-bisphosphate. Hybrid enzymes were produced from the wild-type subunit and a mutant subunit, whose substrate specificity was altered to that of malate dehydrogenase, and separated to analyze the substrate-induced homotropic activation mechanism. Oxamate, a competitive inhibitor of L-lactate dehydrogenase, was used to mimic the substrate-induced activation of the wild-type subunit as "a regulatory subunit." The malate dehydrogenase activity of the mutant subunit as "the catalytic subunit" of the hybrid enzymes was measured, and the activity of the mutant subunit was activated on the addition of oxamate. Thus, we directly observed the inter-subunit homotropic activation transmitted from the wild-type to the mutant subunit. Moreover, "isomeric" hybrid enzymes that have different structural subunit arrangements but identical subunit compositions showed identical kinetic natures. This indicates that the enzyme maintains its subunit symmetry during the allosteric transition.  相似文献   

13.
Disulfide engineering across subunit interfaces provides a means of inhibiting dissociation during unfolding of multimeric enzymes. Two symmetry-related intersubunit disulfide bridges were introduced across the interface of the dimeric enzyme triosephosphate isomerase from Plasmodium falciparum. This was achieved by mutating a tyrosine residue at position 74 at the subunit interface to a cysteine, thereby enabling it to form a covalent cross-link with a pre-existing cysteine at position 13 of the other subunit. The wild-type enzyme (TIMWT) and the oxidized (Y74Cox) and reduced (Y74Cred) forms of the mutant have similar enzymatic activity, absorption, and fluorescence spectra. All three proteins have similar far-UV CD spectra. The Y74Cred shows a distinct loss of near-UV CD. Thermal precipitation studies demonstrate that TIMWT and Y74Cox have very similar Tm values (Tm approximately 60 degreesC) whereas Y74Cred is surprisingly labile (Tm approximately 38 degreesC). The Y74C mutant results in the creation of a large cavity (approximately 100 A3) at the dimer interface. The crystal structure for the oxidized form of Y74C mutant, crystallized in the presence of low concentrations of dithiothreitol, reveals an asymmetric dimer containing a disulfide bridge at one site and a reduced dithiol cysteine at the other. The crystal structure of the mutant offers insights into the destabilization effects of the interfacial cavities and the role of disulfide tethering in restoring protein stability.  相似文献   

14.
The role played by the 6-S-cysteinyl-FMN bond of trimethylamine dehydrogenase in the reductive half-reaction of the enzyme has been studied by following the reaction of the slow substrate diethylmethylamine with a C30A mutant of the enzyme lacking the covalent flavin attachment to the polypeptide. Removal of the 6-S-cysteinyl-FMN bond diminishes the limiting rate for the first of the three observed kinetic phases of the reaction by a factor of 6, but has no effect on the rate constants for the two subsequent kinetic phases. The flavin in the C30A enzyme recovered from the reaction of the C30A enzyme with excess substrate is found to have been converted to the 6-hydroxy derivative, rendering the enzyme inactive. The noncovalently bound FMN of the C30A mutant enzyme is also converted to 6-hydroxy-FMN and rendered inactive upon reduction with excess trimethylamine, but not by reduction with dithionite, even at high pH or in the presence of the effector tetramethylammonium chloride. These results suggest that one significant role of the 6-S-cysteinyl-FMN bond is to prevent the inactivation of the enzyme during catalysis. A reaction mechanism is proposed whereby OH- attacks C-6 of a flavin-substrate covalent adduct in the course of steady-state turnover to form 6-hydroxy-FMN.  相似文献   

15.
Horse liver alcohol dehydrogenase contains two tryptophan residues per subunit, Trp-15 on the surface of the catalytic domain and Trp-314 buried in the interface between the subunits of the dimer. We studied the contributions of the tryptophans to fluorescence and catalytic dynamics by substituting Trp-314 with a leucine residue and making two compensatory mutations that were required to obtain a stable protein, leading to the triple mutant M303F-L308I-W314L enzyme. The substitutions increased by two- to sixfold the turnover numbers for ethanol oxidation, acetaldehyde reduction, and the dissociation constants of the coenzymes. The rate of the exponential burst phase for the transient oxidation of ethanol increased slightly, but the rate of dissociation of the enzyme-NADH complex still limited turnover of ethanol, as for wild-type enzyme. The three substitutions at the dimer interface apparently activate the enzyme by allowing more rapid conformational changes that accompany coenzyme binding, probably due to movement of the loop containing residues 293 to 298. The emission spectrum of M303F-L308I-W314L enzyme, which contains Trp-15, was redshifted compared to wild-type enzyme. Time-resolved fluorescence measurements with the triple mutant show that the decay of Trp-15 is dominated by a approximately 7-ns component. In the mutant enzyme with Trp-15 substituted with phenylalanine, the decay of Trp-314 is dominated by a approximately 4-ns component. Solute quenching data for wild-type enzyme and the mutants show that only Trp-15 is exposed to iodide and acrylamide, whereas Trp-314 is inaccessible. The luminescence properties of the tryptophan residues in the mutated enzymes are consistent with conclusions from studies of the wild-type enzyme [M. R. Eftink, 1992, Adv. Biophys. Chem. 2, 81-114].  相似文献   

16.
We employed site-directed mutagenesis based on sequence comparisons and characterization of purified mutant enzymes to identify Glu558 and Asp766 of Syrian hamster 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase (EC 1.1.1.34) as essential for catalysis. Mutant enzymes E558D, E558Q, and D766N had wild-type Km values for (S)-HMG-CoA and NADPH, but exhibited less than 0.5% of the wild-type catalytic activity. The inactive mutant polypeptides E558Q and D766N nevertheless can associate to generate an active enzyme. In vitro, 6% of the wild-type activity was observed when mutant polypeptides E558D and D766N were mixed in the absence of chaotropic agents. When mutant polypeptides E558Q and D766N were co-expressed in Escherichia coli, the resulting purified enzyme had 25% of wild-type activity. Hamster HMG-CoA reductase thus is a two-site, dimeric enzyme whose subunits associate to form an active site in which each monomer contributes at least one residue (e.g. Glu558 from one monomer and Asp766 from the other). The wild-type enzyme behaves as a dimer during size exclusion chromatography and has one HMG-CoA binding site per monomer. Syrian hamster HMG-CoA reductase thus appears to be a homodimer with two active sites which are located at the subunit interface.  相似文献   

17.
The conserved residue Asp477 in yeast transketolase is located in the substrate channel of the enzyme and forms a hydrogen bond with the C2-hydroxyl group of the acceptor substrate. The significance of this interaction for the recognition of the preferred acceptor substrates, D-alpha-hydroxyaldehydes was investigated by site-directed mutagenesis. In the wild-type enzyme the kcat/KM values are by three to four orders of magnitude lower for 2-deoxyaldoses or substrates with L-configuration at the C2-atom. In the Asp477 Ala mutant, the kcat/KM values for D-alpha-hydroxyaldehydes are decreased by a thousandfold, while the kcat/KM values for substrates with L-configuration or 2-deoxyaldoses are similar to wild-type enzyme. These results indicate that Asp477 is involved in determining the enantioselectivity of transketolase.  相似文献   

18.
To generate novel forms of metal-binding proteins, six mutant mouse metallothionein (MT) 1 fragments, in which a terminal cysteine residue was replaced by histidine, were expressed in Escherichia coli. The spectroscopic and analytical results showed that the alphaMT (C33H, C36H, C41H, C57H) and betaMT (C5H, C13H) mutant forms bound 4 and 3 Zn(II) atoms per molecule of protein to the nearest integer, even though in C41H and C5H, species of lower stoichiometry were also detected. In Cd(II) titrations, all the Zn(II) ions bound to the mutant proteins were displaced from the binding sites, giving rise to Cd-mutated MT forms with 4 and 3 Cd(II), respectively. However, although Cys-to-His substitutions maintained the binding capacity of the MT fragments, they caused structural changes with respect to the wild-type proteins. While C13H, C36H and C57H seem to contain Zn(II)-aggregates that are closely related to those of the wild-type proteins, only C41H and C57H gave rise to Cd(II)-aggregates similar to those of Cd4-alphaMT, where the His residue plays the role of the substituted Cys. Despite the structural implications of the Cys-to-His replacement, the dissociation constants showed no major decrease in the Cd-binding affinity in any of the mutants assayed compared with the wild-type.  相似文献   

19.
OBJECTIVE: To characterize the biochemical mechanisms of expression of the pyruvate dehydrogenase (PDH) E1alpha subunit exon 10 R302C missense mutation. BACKGROUND: Mutations in the X-linked E1alpha subunit gene are responsible for most cases of PDH deficiency, an important cause of neurodevelopmental defects and neurodegeneration with primary lactic acidemia. Although the disease shows extreme allelic heterogeneity, the R302C mutation has been defined in several unrelated cases. METHODS: Cell lines expressing selectively either the mutant or wild-type E1alpha alleles against identical genetic backgrounds were generated from the fibroblasts of a female heterozygous for the R302C mutation. Enzyme activity, mRNA, polypeptide expression, and turnover were studied in each. RESULTS: The residual PDH activity was below measurable levels in the cell line (B5) expressing only the mutant allele and normal in the wild-type polypeptide expressing (A10) cell line, confirming that the R302C mutation alone is sufficient to cause a severe PDH deficiency. The mutant polypeptide was less stable than the wild-type polypeptide, but the steady-state level of the mutant E1alpha protein was reduced only two- to threefold. CONCLUSIONS: The primary mechanism of expression of the R302C mutation must be limitation of catalytic efficiency. We speculate that catalysis may be inhibited in the mutant polypeptide because conformational changes are induced near serine 300, a residue that is particularly important as a regulatory phosphorylation site in the wild-type polypeptide.  相似文献   

20.
Xanthine dehydrogenase, a molybdenum, iron-sulfur flavoenzyme encoded in the fruit fly Drosophila melanogaster by the rosy gene, has been characterised both from the wild-type and mutant files. Enzyme assays, using a variety of different oxidising and reducing substrates were supplemented by limited molecular characterisation. Four rosy strains showed no detectable activity in any enzyme assay tried, whereas from four wild-type and three rosy mutant strains, those for the [E89K], [L127F] and [L157P]xanthine dehydrogenases (in all of which the mutation is in the iron-sulfur domain), the enzyme molecules, although present at different levels, had extremely similar or identical properties. This was confirmed by purification of one wild-type and one mutant enzyme. [E89K]xanthine dehydrogenase. These both had ultraviolet-visible absorption spectra similar to milk xanthine oxidase. Both were found to be quite stable molecules, showing very high catalytic-centre activities and with little tendency to become degraded by proteolysis or modified by conversion to oxidase or desulfo forms. In three further rosy strains, giving [G353D]xanthine dehydrogenase and [S357F]xanthine dehydrogenase mutated in the flavin domain, and [G1011E]xanthine dehydrogenase mutated in the molybdenum domain, enzyme activities were selectively diminished in certain assays. For the G353D and S357F mutant enzymes activities to NAD+ as oxidising substrate were diminished, to zero for the latter. In addition for [G353D]xanthine dehydrogenase, there was an increase in apparent Km values both for NAD+ and NADH. These findings indicate involvement of this part of the sequence in the NAD(+)-binding site. The G1011E mutation has a profound effect on the enzyme. As isolated and as present in crude extracts of the files, this xanthine dehydrogenase variant lacks activity to xanthine or pterin as reducing substrate, indicating an impairment of the functioning of its molybdenum centre. However, it retains full activity to NADH with dyes as oxidising substrate. Mild oxidation of the enzyme converts it, apparently irreversibly, to a form showing full activity to xanthine and pterin. The nature of the group that is oxidised is discussed in the light of redox potential data. It is proposed that the process involves oxidation of the pterin of the molybdenum cofactor from the tetrahydro to a dihydro oxidation state. This conclusion is fully consistent with recent information [Rom?o, M. J., Archer, M., Moura, I., Moura. J.J.G., LeGall, J., Engh, R., Schneider, M., Hof, P. & Huber, R. (1995) Science 270. 1170-1176) from X-ray crystallography on the structure of a closely related enzyme from Desulfovibrio gigas. It is proposed, that apparent irreversibility of the oxidative activating process for [G1011E]xanthine dehydrogenase, is due to conversion of its pterin to the tricyclic derivative detected by these workers. The data thus provide the strongest evidence available, that the oxidation state of the pterin can have a controlling influence on the activity of a molybdenum cofactor enzyme. Implications regarding pterin incorporation into xanthine dehydrogenase and in relation to other molybdenum enzymes are discussed.  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