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 49 毫秒
1.
Reaction of the NADP-dependent glutamate dehydrogenase of Neurospora with 1,2-cyclohexanedione results in a biphasic loss of enzyme activity. At the end of the rapid phase of the reaction (t1/2 = 1.5 min) the enzyme activity is diminished by approximately 60% with the simultaneous loss of 1 residue of arginine per subunit. After 60 min, the enzyme activity is completely lost with the modification of a total of 2 arginine residues per subunit. Reaction of bovine liver glutamate dehydrogenase with cyclohexanedione causes a rapid loss of approximately 45% of the enzyme activity and modification of about 1.5 residues of arginine per subunit. More prolonged treatment results in reaction of an additional 4 residues of arginine per subunit but is without further effect on the residual activity. The activity of the Neurospora enzyme is not protected by substrate, coenzyme, or a combination of both; however, the activity of the bovine enzyme is partially protected by high levels of NAD or NADP. Although the Km for alpha-ketoglutarate is unchanged by a limited modification of either enzyme with cyclohexanedione, the Km for coenzyme is increased about 2-fold for the Neurospora enzyme and about 1.5-fold for the bovine enzyme. The Ki of the Neurospora dehydrogenase for the competitive inhibitor 2'-monophosphoadenosine-5'-diphosphoribose is unchanged by the enzyme modification, but nicotinamide mononucleotide, a competitive inhibitor for the native Neurospora enzyme, does not inhibit the glutamate dehydrogenase with 1 modified arginine residue. This finding implies that the modified arginine is at or near the nicotinamide binding iste of the enzyme.  相似文献   

2.
A sequence is presented for the COOH-terminal 669 residues of the NAD-specific glutamate dehydrogenase of Neurospora crassa. Comparison of this sequence with those of the vertebrate glutamate dehydrogenases of chicken and bovine liver and with the NADP-specific enzyme of Neurospora shows some similarities in sequences around residues previously identified as important for the function of these enzymes. These are: (a) the reactive lysine residue of low pK in the NADP and the vertebrate enzymes; (b) the tyrosine residue of the NADP enzyme that is readily nitrated by tetranitromethane with inactivation, a residue protected by NADP or by NMN; and (c) the arginine residue of the NADP-enzyme that is reactive with 1,2-cyclohexanedione with inactivation. Despite these similarities, comparison of the sequence of the NAD-enzyme with those of the other glutamate dehydrogenases of known sequences revealed relatively little overall homology as determined by computer analysis.  相似文献   

3.
The C-terminal domain (C(t)-FDH) of 10-formyltetrahydrofolate dehydrogenase (FDH, ALDH1L1) is an NADP(+)-dependent oxidoreductase and a structural and functional homolog of aldehyde dehydrogenases. Here we report the crystal structures of several C(t)-FDH mutants in which two essential catalytic residues adjacent to the nicotinamide ring of bound NADP(+), Cys-707 and Glu-673, were replaced separately or simultaneously. The replacement of the glutamate with an alanine causes irreversible binding of the coenzyme without any noticeable conformational changes in the vicinity of the nicotinamide ring. Additional replacement of cysteine 707 with an alanine (E673A/C707A double mutant) did not affect this irreversible binding indicating that the lack of the glutamate is solely responsible for the enhanced interaction between the enzyme and the coenzyme. The substitution of the cysteine with an alanine did not affect binding of NADP(+) but resulted in the enzyme lacking the ability to differentiate between the oxidized and reduced coenzyme: unlike the wild-type C(t)-FDH/NADPH complex, in the C707A mutant the position of NADPH is identical to the position of NADP(+) with the nicotinamide ring well ordered within the catalytic center. Thus, whereas the glutamate restricts the affinity for the coenzyme, the cysteine is the sensor of the coenzyme redox state. These conclusions were confirmed by coenzyme binding experiments. Our study further suggests that the binding of the coenzyme is additionally controlled by a long-range communication between the catalytic center and the coenzyme-binding domain and points toward an α-helix involved in the adenine moiety binding as a participant of this communication.  相似文献   

4.
Initial-rate studies were made of the oxidation of L-glutamate by NAD+ and NADP+ catalysed by highly purified preparations of dogfish liver glutamate dehydrogenase. With NAD+ as coenzyme the kinetics show the same features of coenzyme activation as seen with the bovine liver enzyme [Engel & Dalziel (1969) Biochem. J. 115, 621--631]. With NADP+ as coenzyme, initial rates are much slower than with NAD+, and Lineweaver--Burk plots are linear over extended ranges of substrate and coenzyme concentration. Stopped-flow studies with NADP+ as coenzyme give no evidence for the accumulation of significant concentrations of NADPH-containing complexes with the enzyme in the steady state. Protection studies against inactivation by pyridoxal 5'-phosphate indicate that NAD+ and NADP+ give the same degree of protection in the presence of sodium glutarate. The results are used to deduce information about the mechanism of glutamate oxidation by the enzyme. Initial-rate studies of the reductive amination of 2-oxoglutarate by NADH and NADPH catalysed by dogfish liver glutamate dehydrogenase showed that the kinetic features of the reaction are very similar with both coenzymes, but reactions with NADH are much faster. The data show that a number of possible mechanisms for the reaction may be discarded, including the compulsory mechanism (previously proposed for the enzyme) in which the sequence of binding is NAD(P)H, NH4+ and 2-oxoglutarate. The kinetic data suggest either a rapid-equilibrium random mechanism or the compulsory mechanism with the binding sequence NH4+, NAD(P)H, 2-oxoglutarate. However, binding studies and protection studies indicate that coenzyme and 2-oxoglutarate do bind to the free enzyme.  相似文献   

5.
Conformational changes induced by binding of ligands to cytosolic NADP(+)-specific isocitrate dehydrogenase from lactating bovine mammary gland were assessed using circular dichroism and fluorescence techniques. The secondary structure of isocitrate dehydrogenase, as monitored by CD spectra in the far-UV region, is unaltered by enzyme-ligand interactions; in contrast, dramatic changes occur in the near-UV region (270-290 nm) assigned to tyrosine and/or solvent-exposed tryptophan residues. Both the coenzyme analog, 2'-phosphoadenosine 5'-diphosphoribose, and NADPH have an effect on the CD spectrum which is opposite to that produced by metal complexes of either isocitrate or citrate. A CD band at 292 nm assigned to approximately 2 tryptophan residues in a hydrophobic environment is unchanged by binding of substrate or coenzyme. Approximately 30% of the intrinsic fluorescence of isocitrate dehydrogenase, corresponding to approximately 2 tryptophan residues, is not quenched by acrylamide in the absence of 6.3 M guanidine hydrochloride and remains unquenched in the enzyme-substrate complex. The constancy in the proportion of buried and exposed tryptophan residues implicates tyrosine in the observed near-UV CD spectral changes. Since binding of ligands does not influence quaternary structure (Seery, V.L., and Farrell, H. M., Jr. (1989) Arch. Biochem. Biophys. 274, 453-462), activation of isocitrate dehydrogenase may be related to a substrate-induced conformational transition.  相似文献   

6.
Extracts of Salmonella typhimurium were chromatographed by using Sephadex G-150 to separate the various enzymes involved with pyridine nucleotide cycle metabolism. This procedure revealed a previously unsuspected nicotinamide adenine dinucleotide (NAD) glycohydrolase (EC 3.2.2.5) activity, which was not observed in crude extracts. In contrast to NAd glycohydrolase, NAD pyrophosphatase (EC 3.6.1.22) was readily measured in crude extracts. This enzyme possessed a native molecular weight of 120,000. Other enzymes examined included nicotinamide mononucleotide (NMN) deamidase (EC 3.5.1.00), molecular weight of 43,000; NMN glycohydrolase (EC 3.2.2.14), molecular weight of 67,000; nicotinic acid phosphoribosyl transferase (EC 2.4.2.11), molecular weight of 47,000; and nicotinamide deamidase (EC 3.5.1.19), molecular weight of 35,000. NMN deamidase and NMN glycohydrolase activities were both examined for end product repression by measuring their activities in crude extracts prepared from cells grown with and without 10(-5) M nicotinic acid. No repression was observed with either activity. Both activities were also examined for feedback inhibition by NAD, reduced NAD, and NADP. NMN deamidase was unaffected by any of the compounds tested. NMN glycohydrolase was greatly inhibited by NAD and reduced NAD, whereas NADP was much less effective. Inhibition of NMN glycohydrolase was found to level off at an NAD concentration of ca. 1 mN, the approximate intracellular concentration of NAD.  相似文献   

7.
Phenylacetaldehyde dehydrogenase (PAD) and lactaldehyde dehydrogenase (ALD) share some structural and kinetic properties. One difference is that PAD can use NAD+ and NADP+, whereas ALD only uses NAD+. An acidic residue has been involved in the exclusion of NADP+ from the active site in pyridine nucleotide-dependent dehydrogenases. However, other factors may participate in NADP+ exclusion. In the present work, analysis of the sequence of the region involved in coenzyme binding showed that residue F180 of ALD might participate in coenzyme specificity. Interestingly, F180T mutation rendered an enzyme (ALD-F180T) with the ability to use NADP+. This enzyme showed an activity of 0.87 micromol/(min * mg) and K(m) for NADP+ of 78 microM. Furthermore, ALD-F180T exhibited a 16-fold increase in the V(m) /K(m) ratio with NAD+ as the coenzyme, from 12.8 to 211. This increase in catalytic efficiency was due to a diminution in K(m) for NAD+ from 47 to 7 microM and a higher V(m) from 0.51 to 1.48 micromol/(min * mg). In addition, an increased K(d) for NADH from 175 (wild-type) to 460 microM (mutant) indicates a faster product release and possibly a change in the rate-limiting step. For wild-type ALD it is described that the rate-limiting step is shared between deacylation and coenzyme dissociation. In contrast, in the present report the rate-limiting step in ALD-F180T was determined to be exclusively deacylation. In conclusion, residue F180 participates in the exclusion of NADP+ from the coenzyme binding site and disturbs the binding of NAD+.  相似文献   

8.
A glycohydrolase that catalyzes the irreversible conversion of NMN to nicotinamide and ribose 5-phosphate has been partially purified from a sonic extract of Azotobacter vinelandii. The enzyme is highly specific for NMN. NAD, NADP, nicotinic acid-adenine dinucleotide, nicotinamide riboside and alpha-NMN are not significantly hydrolyzed by this enzyme, nor do they compete with NMN. The enzyme also exhibits an absolute dependence on guanylic acid derivatives with following order of relative effectiveness: GTP, guanosine 5'-tetraphosphate greater than dGTP, GDP, 2'-GMP, 3'-GMP greater than GMP, dGMP. A heat-resistant, nondialyzable factor which could replace the GTP requirement was found in the sonic extract. The Ka for GTP and the Km for NMN in the presence of GTP at 1mm were calculated to be 0.025 mM and 4.5 mM respectively. GMP, dGMP, and dCMP were found to be effective inhibitors of the enzyme when 1 mM GTP was also present. The kinetic data suggest that the binding site for these mononucleotides is distinct from the active site or the GTP binding site. The ability of this enzyme to cleave NMN is suggestive of a metabolic role of the enzyme in selective conversion of NMN to nicotinamide, which, in turn, would be re-utilized by the cell as a precursor of NAD via nicotinic acid.  相似文献   

9.
The crystal structure of sheep liver 6-phosphogluconate dehydrogenase (6PGDH) shows marked differences in the position of the nicotinamide mononucleotide (NMN) moiety of NADP(+) and NADPH (Adams, J. M., Grant, H. E., Gover, S., Naylor, C. E., and Phillips, C. (1994) Structure 2, 651-668). A methionine side chain (Met13) interacts with the si face of NADP(+) in the complex with the oxidized coenzyme, is likely to affect the binding mode of the nicotinamide ring of NADP(+), and may play a role in catalysis in the 6PGDH reaction. To check this possibility we performed site-directed mutagenesis, changing M13 to a number of residues including V, I, C, F, and Q. Mutant enzymes were characterized with respect to their kinetic parameters and primary deuterium isotope effects. All mutations resulted in a decrease in affinity of the enzyme for NADP(+), but not NADPH. In addition, the M13 to C (M13C), M13F, and M13Q mutant enzymes exhibited a decrease of at least an order of magnitude in V/E(t). The deuterium isotope effects on V and V/K(6PG) were decreased to about 1.2 for the M13F and M13C mutant enzymes, while they were increased to about 2.4 for the M13Q enzyme (a value of 1.8-1.9 is obtained for the wild-type enzyme). In at least three instances changes in the overall rate of the oxidative decarboxylation reaction relative to other steps along the reaction pathway were observed. Isotope effects indicate that the hydride transfer steps can become either more or less rate-determining dependent on the substitution. Data are consistent with a significant role of M13 in the orientation of the cofactor nicotinamide ring in the mechanism of 6PGDH, likely with respect to geometry and distance of the ring from C3 of 6PG.  相似文献   

10.
The dinucleotide binding beta alpha beta motif in the crystal structures of seven different enzymes has been analysed in terms of their three-dimensional structures and primary sequences. We have identified that the hydrogen bonding of the adenine ribose to the glycine-rich turn containing the fingerprint sequence GXGXXG/A occurs via a direct or indirect mechanism, depending on the nature of the fingerprint sequence but independent of coenzyme specificity. The major determinant of the type of interaction is the nature of the residue occupying the last position of the above fingerprint. In the NAD(+)-linked dehydrogenases, an acidic residue is commonly used to form important hydrogen bonds to the adenine ribose hydroxyls and, hitherto, this residue has been thought to be an indicator of NAD+ specificity. However, on the basis of the three-dimensional structure of the NAD(+)-linked glutamate dehydrogenase (GDH) from Clostridium symbiosum we have demonstrated that this residue is not a universal requirement for the construction of an NAD+ binding site. Furthermore, considerations of sequence homology unambiguously identify an equivalent acidic residue in both NADP+ and dual specificity glutamate dehydrogenases. The conservation of this residue in these enzymes, coupled to its close proximity to the 2' phosphate implied by the necessary similarity in three-dimensional structure to C. symbiosum GDH, implicates this residue in the recognition of the 2' phosphate either via water-mediated or direct hydrogen-bonding schemes. Analysis of the latter has led us to suggest that two patterns of recognition for the 2' phosphate group of NADP(+)-binding enzymes may exist, which are distinguished by the ionization state of the 2' phosphate.  相似文献   

11.
6-Phosphogluconate dehydrogenase is the pivotal enzyme that links the gluconate route and the oxidative phase of the pentose phosphate pathway in Schizosaccharomyces pombe. The enzyme differs from the known 6-phosphogluconate dehydrogenases of other sources in that the Schizosaccharomyces enzyme is tetrameric having a subunit mass of 38 kDa, that it requires NADP+ obligatorily for activity, and that it can be activated by divalent metal ions such as Co2+ and Mn2+. Steady-state kinetic studies were undertaken. Initial rate and product inhibition results suggest that 6-phosphogluconate dehydrogenase from Schizosaccharomyces pombe catalyzes NADP(+)-linked oxidative decarboxylation of 6-phosphogluconate by an equilibrium random mechanism with two independent binding sites, namely one site for the nicotinamide coenzyme, NADP+/NADPH, and another site for 6-phosphogluconate-D-ribulose-5-phosphate and for CO2. Studies of pH dependence implicated a basic residue with a pK value of 7.4 in the binding of 6-phosphogluconate and an acidic residue with a pK value of 6.7 in the cation-mediated interaction of NADP+ with the enzyme.  相似文献   

12.
Structural and conformational organization of chicken liver fatty acid synthetase has been probed using its fluorescent coenzyme, NADPH. Three NADPH binding sites per mole of the enzyme complex, of apparently identical dissociation constant (KD = 0.6 muM) can be titrated at temperatures above 12 degrees. These results are in disagreement with the earlier studies of Hsu and Wagner (Hsu, R. Y., and Wagner, B. J. (1970) Biochemistry, 9, 245-251) in which four such sites could be titrated. At 12 degrees, the composite sites split into two subsets: a pair of sites with a KD of 0.3 muM and a third site with a Kd of 1.1 muM. At lower temperatures (5 degrees or 2 degrees), the site with weak affinity disappears, leaving a pair of sites with a Kd of 0.5 muM. Similar observations were made when the enzyme was modified with phenylmethylsulfonyl fluoride, a specific and selective inhibitor of fatty acyl-CoA deacylase (s) of the pigeon liver enzyme complex (Kumar, S. (1975) J. Biol. Chem. 250, 5150-5158). Partial modification with phenylmethylsulfonyl fluoride elicits a NADPH binding response similar to the binding observed at 12 degrees, i.e. two sets of binding sites with nonidentical dissociation constants. Further modification corresponding to the complete loss of deacylase function results in a set of two apparently identical binding sites, and the third site is not available for titration. The modified enzyme retains the two reductase functions as measured by the model substrates, acetoacetyl-N-acetylcysteamine and crotonyl-CoA. Furthermore, the addition of acetyl- and malonyl-CoA (100 muM each) to the modified enzyme lowers the NADPH binding affinity by a factor of 3. Other observations show that the quantum yield, as measured by the ratio of fluorescence intensity of bound and free NADPH, changes with temperature and ionic strength. Lowering the temperature from 30 degrees to 2 degrees increases the enhancement ratio by 50%, whereas increase in ionic strength from 0.05 to 0.2 M potassium phosphate lowers it to 50% of the original level. Measurement of NADPH binding in the presence of NADP+, NADH, NAD+ and adenosine-2'-monophospho-5'-diphosphoribose demonstrates that NADP+ shows competitive behavior for NADPH sites (KD = 10.6 muM), whereas NADH and NAD+ show noncompetitive (KD (apparent) = nearly 600 muM) and rather complicated interactions implicating nonspecific conformational alteration of the enzyme complex. The behavior of adenosine 2'-monophospho-5'-diphosphoribose is intermediate between NADP+ and NADH. These data are discussed in terms of substrate-mediated conformational changes and the moles of each of the reductase enzymes per mole of the enzyme complex, the polarity of the NADPH binding region, and the probable structure of the nicotinamide moiety when bound to the enzyme.  相似文献   

13.
Fatty acid synthase from the uropygial gland of goose was inhibited by treatment with pyridoxal 5'-phosphate by selectively modifying a lysine residue at the NADPH binding site of the enoyl reductase domain (A. J. Poulose and P. E. Kolattukudy (1980) Arch. Biochem. Biophys. 201, 313-321). Distribution of radioactivity in tryptic peptides generated from the synthase treated with pyridoxal 5'-phosphate/NaB3H4 in the presence and absence of 2'-monophosphoadenosine-5'-diphosphoribose, which protects the enzyme from inactivation by pyridoxal phosphate, showed that modification of one specific peptide was prevented by the protector. This peptide was purified by a combination of Sephadex G-25 column chromatography, anion-exchange chromatography, and high-performance liquid chromatography. The primary structure of this peptide is Val-Phe-Thr-Thr-Val-Gly-Ser-Ala-Glu-Lys(Pxy)-Arg.  相似文献   

14.
The mechanism of action of yeast beta-hydroxy-beta-methylglutaryl-coenzyme A reductase has been investigated through kinetic studies on the oxidation of mevaldate by nicotinamide adeninine dinucleotide phosphate (NADP) in the presence of coenzyme A (CoA) and on the reduction of mevaldate by reduced NADP (NADPH) in the absence of presence of CoA or acetyl-CoA. NADP and mevalonate were also used as product inhibitors of the reduction of mevaldate. In the reduction of mevaldate to mevalonate, coenzyme A and acetyl-CoA decreased the Km for mevaldate 30- and 3-fold, respectively. Both compounds increased the Vmax 1.5-fold. These results suggest that CoA is an allosteric activator for the second reductive step and that it acts by enhancing the binding of mevaldate. The intersecting patterns obtained from initial velocities and the patterns produced by product inhibitions suggest the following features of the mechanism. The binding of substrates and release of products proceeds sequentially in both reductive steps, and is ordered throughout or random with respect to the binding of the beta-hydroxy-beta-methylglutaryl-coenzymeA and the first NADPH. The binding of NADPH enhances the binding of the beta-hydroxy-beta-methylglutaryl portion of the CoA ester and the binding of free mevaldate, whereas the binding of NADP leads to an increased affinity of the enzyme for the hemithioacetal (of mevaldate and CoA) and for mevalonate. Thus, the replacement of NADP by NADPH after the first reductive step promotes the conversion of the hemithioacetal to the free carbonyl form, which is then rapidly reduced. The products, CoA and mevalonic acid, of the second reductive step leave the enzyme before the release of the second NADP. This release of the last product is probably the rate-limiting step for the overall process.  相似文献   

15.
The structural requirements of the NADP+ molecule as a coenzyme in the oxidative decarboxylation reaction catalysed by pigeon liver malic enzyme were studied by kinetic and fluorimetric analyses with various NADP+ analogues and fragments. The substrate L-malate had little effect on the nucleotide binding. Etheno-NADP+, 3-acetylpyridine-adenine dinucleotide phosphate, and nicotinamide-hypoxanthine dinucleotide phosphate act as alternative coenzymes for the enzyme. Their kinetic parameters were similar to that of NADP+. Thionicotinamide-adenine dinucleotide phosphate, 3-aminopyridine-adenine dinucleotide phosphate, 5'-adenylyl imidodiphosphate, nicotinamide-adenine dinucleotide 3'-phosphate and NAD+ act as inhibitors for the enzyme. The first two were competitive with respect to NADP+ and non-competitive with respect to L-malate; the other inhibitors were non-competitive with NADP+. All NADP+ fragments were inhibitory to the enzyme, with a wide range of affinity, depending on the presence or absence of a 2'-phosphate group. Compounds with this group bind to the enzyme 2-3 orders of magnitude more tightly than those without this group. Only compounds with this group were competitive inhibitors with respect to NADP+. We conclude that the 2'-phosphate group is crucial for the nucleotide binding of this enzyme, whereas the carboxyamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity. There is a strong synergistic effect between the binding of the nicotinamide and adenosine moieties of the nucleotide molecule.  相似文献   

16.
Rat liver S-adenosylhomocysteinase, a homotetramer, was resolved by treatment with acid ammonium sulfate into apoenzyme and NAD. The apoenzyme thus prepared retained a tetrameric structure but differed in the mobility on nondenaturing polyacrylamide gel electrophoresis. The inactive apoenzyme was reactivated upon incubation with NAD. The restoration of activity paralleled with the tight binding of NAD to apoenzyme, and full activity was obtained when 4 mol of NAD were bound per mol of apoenzyme. The kinetics of reconstitution were apparently biphasic and suggest the existence of two conformers in a slow equilibrium, one of which binds the coenzyme rapidly while the other does so very slowly, if at all. In addition to NAD, apoadenosylhomocysteinase tightly bound nicotinamide hypoxanthine dinucleotide, 3-acetylpyridine adenine dinucleotide and nicotinic acid-adenine dinucleotide. NADP was not bound. Catalytic activity was found only with the enzyme reconstituted with NAD or nicotinamide hypoxanthine dinucleotide. The spectral change observed on interaction of apoadenosylhomocysteinase with NAD was similar to those seen with adenine nucleotides, and was largely approximated by the addition of dioxane to aqueous solutions of adenine nucleotides. By comparison of the difference spectra, it is suggested that the adenine portion of the coenzyme is bound in the hydrophobic pocket of the protein, and that the binding is accompanied by perturbation of tryptophan residue of the protein.  相似文献   

17.
The oxidized coenzyme NAD binds to two sites per subunit of bovine liver glutamate dehydrogenase with equal affinity in the absence of dicarboxylic acid coligands. In the presence of glutarate or 2-oxoglutarate, the affinity to one site is unchanged, but the affinity to the other (presumed to be the active site) is considerably increased and now requires two dissociation constants to describe its saturation. A combination of transfer nuclear Overhauser effects (TRNOE) together with an examination of the slopes of TRNOE time dependence indicates that while NAD is bound in a syn conformation at both binding sites, NADP (which binds only to the active site) is bound in a syn-anti mixture. The existence of N6 to N3' and N6 and N2' and N1' to N3' NOE's with NAD suggests that the two coenzyme binding sites are located near enough to allow intermolecular NOE's. In the presence of 2-oxoglutarate where only binding to the active site is effectively observed, the conformation of either coenzyme is syn. Modeling studies using the distance estimates from the TRNOE results suggest that the nicotinamide ribose approximates a 3'-endo conformation. The absence of evidence for intermolecular NOE's under these conditions indicates that while the active and regulatory NAD sites per subunit are in close proximity, the six active sites per hexamer are located greater than 5 A apart.  相似文献   

18.
Glutamate dehydrogenase (GDH) (L-glutamate:NADP+ oxidoreductase, deaminating, EC 1.4.1.4) from the cellulolytic ruminal bacterium Ruminococcus flavefaciens has been purified and characterized. The native enzyme and subunit are 280 and 48 kDa, respectively, suggesting that the native enzyme is a hexamer. The enzyme requires 0.5 M KCl for optimal activity and has a pH optimum of 6.9 to 7.0. The Kms for ammonia, alpha-ketoglutarate, and glutamate are 19, 0.41, and 62 mM, respectively. The sigmoidal NADPH saturation curve revealed positive cooperativity for the binding of this coenzyme. The first residue in the N-terminal amino acid sequence from R. flavefaciens GDH was alanine, suggesting that the protein may be modified posttranslationally. Comparison of the N-terminal sequence with those of Escherichia coli, Salmonella typhimurium, and Clostridium symbiosum revealed only 39% amino acid homologies. The GDH from R. flavefaciens was unique in that its specific activity was highest during ammonia-limited growth but was not affected by ammonia shock treatment (20 mM).  相似文献   

19.
The NADP-specific glutamate dehydrogenase (EC 1.4.1.4) of Neurospora crassa is inhibited by reaction with 1,2-cyclohexanedione which binds to arginine residues. With the 14C-labeled reagent, a peptide was isolated with the sequence: Gly-Gly-Leu-Arg-Leu-His-Pro-Ser-Val-Asn-Leu, corresponding to residues 78 through 88 in the protein. The arginine, residue 81, was present as N7,N8-(1,2-dihydroxycyclohex-1,2-ylene)-arginyl (or DHCH-arginine). Present evidence indicates that this arginine residue resides at or near the nicotinamide binding domain of the enzyme. Similar sequences are present in the bovine liver enzyme (EC 1.4.1.3) and the NAD-specific glutamate dehydrogenase of Neurospora (EC 1.4.1.2).  相似文献   

20.
The reaction of iodine with glyceraldehyde 3-phosphate dehydrogenase from Bacillus stearothermophilus was investigated. The active-site thiol group of the cysteine residue homologous with cysteine-149 in the pig muscle enzyme was protected by reaction with tetrathionate. The apoenzyme was readily inhibited by KI3 solution at pH8, but the coenzyme, NAD+, protected the enzyme against inhibition and decreased the extent of iodination. At pH 9.5, ready inhibition of both apo- and holo-enzyme was observed. Tryptic peptides containing residues iodinated at pH 8 were isolated and characterized. One of the most reactive residues in both holo- and apo-enzymes was a tyrosine homologous with tyrosine-46 in the pig muscle enzyme, and this residue was iodinated without loss of enzymic activity. Other reactive tyrosine residues in the apoenzyme were in positions homologous with residues 178, 273, 283 and 311 in the pig muscle enzyme, but they were not readily iodinated in the holoenzyme. Histidine residues in both holo- and apo-enzymes were iodinated at pH 8 in sequence positions homologous with residues 50, 162 and 190 in the pig muscle enzyme. The inhibition of the enzyme was not correlated with the iodination of a particular residue. The results are discussed in relation to a three-dimensional model based on the structure of the lobster muscle enzyme and demonstrate that conformational changes affecting the reactivity of several tyrosine residues most probably occur on binding of the coenzyme.  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