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.
Aspartic acid 244 that occurs at the putative NAD(+)-binding site of rat liver S-adenosylhomocysteinase was replaced by glutamic acid by oligonucleotide-directed mutagenesis. The mutant enzyme was purified to homogeneity as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Gel permeation chromatography showed that the purified mutant enzyme was a tetramer as is the wild-type enzyme. In contrast to the wild-type enzyme, which possesses 1 mol of tightly bound NAD+ per mol of enzyme subunit, the mutant enzyme had only 0.05 mol of NAD+ but contained about 0.6 mol each of NADH and adenine per mol of subunit. The mutant enzyme, after removal of the bound compounds by acid-ammonium sulfate treatment, exhibited S-adenosylhomocysteinase activity when assayed in the presence of NAD+. From the appearance of activity as a function of NAD+ concentration, the enzyme was shown to bind NAD+ with a Kd of 23.0 microM at 25 degrees C, a value greater than 280-fold greater than that of the wild-type enzyme. In the presence of a saturating concentration of NAD+, the mutant enzyme showed apparent Km values for substrates similar to those of the wild-type enzyme. Moderate decreases of 8- and 15-fold were observed in Vmax values for the synthetic and hydrolytic directions, respectively. These results indicate the importance of Asp-244 in binding NAD+, and are consistent with the idea that the region of S-adenosylhomocysteinase from residues 213 to 244 is part of the NAD+ binding site. This region has structural features characteristic of the dinucleotide-binding domains of NAD(+)- and FAD-binding proteins (Ogawa, H., Gomi, T., Mueckler, M. M., Fujioka, M., Backlund, P.S., Jr., Aksamit, R.R., Unson, C.G., and Cantoni, G.L. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 719-723).  相似文献   

2.
C K Barlowe  D R Appling 《Biochemistry》1990,29(30):7089-7094
An NAD(+)-dependent 5,10-methylenetetrahydrofolate (THF) dehydrogenase has been purified to homogeneity from the yeast Saccharomyces cerevisiae. The purified enzyme exhibits a final specific activity of 5.4 units mg-1 and is represented by a single protein of apparent Mr = 33,000-38,000 as determined by sodium dodecyl sulfate gel electrophoresis. A native Mr = 64,000 was determined by gel filtration, suggesting a homodimer subunit structure. Cross-linking experiments with dimethyl suberimidate confirmed the dimeric structure. The enzyme is specific for NAD+ and is not dependent on Mg2+ for activity. The forward reaction initial velocity kinetics are consistent with a sequential reaction mechanism. With this model, Km values for NAD+ and (6R,S)-5,10-methylene-THF are 1.6 and 0.06 mM, respectively. In contrast to all other previously described eukaryotic 5,10-methylene-THF dehydrogenases, the purified enzyme is apparently monofunctional, with undetectable 5,10-methenyl-THF cyclohydrolase and 10-formyl-THF synthetase activities. Subcellular fractionation of yeast indicates the enzyme is cytoplasmic, with no NAD(+)-dependent 5,10-methylene-THF dehydrogenase detectable in mitochondria. The activity was found in all yeast strains examined, at all stages of growth from the lag phase through the stationary phase.  相似文献   

3.
Human liver extracts show two major bands with aldehyde dehydrogenase (Aldehyde:NAD+ oxidoreductase, EC 1.2.1.3) activity via starch gel electrophoresis at pH 7.0. Both bands have been purified to apparent homogeneity via classical chromatography combined with affinity chromatography on 5'-AMP-Sepharose 4B. The slower migrating band, enzyme 1, when assayed at pH 9.5 has a low Km for NAD (8 micrometer) and a high Km for acetaldehyde (approx. 0.1 mM). It is very strongly inhibited by disulfiram at pH 7.0 with a Ki of 0.2 micrometer. The faster migrating band, enzyme 2, has a low Km for acetaldehyde, (2--3 micrometer at pH 9.5), a higher Km for NAD (70 micrometer at pH 9.5), and is not inhibited by disulfiram at pH 7.0. The two enzymes are very similar to the F1 and F2 isozymes of horse liver purified by Eckfeldt et al. (Eckfeldt, J., Mope, L., Takio, K. and Yonetani, T. (1976) J. Biol, Chem. 251, 236-240) in molecular weight, subunit composition, amino acid composition and extinction coefficient. Preliminary kinetic characterizations of the enzyme are presented.  相似文献   

4.
The binding of NAD(+) and NADH to bovine liver UDP-glucose dehydrogenase was studied by using gel-filtration and fluorescence-titration methods. The enzyme bound 0.5mol of NAD(+) and 2 mol of NADH/mol of subunit at saturating concentrations of both substrate and product. The dissociation constant for NADH was 4.3mum. The binding of NAD(+) to the enzyme resulted in a small quench of protein fluorescence whereas the binding of NADH resulted in a much larger (60-70%) quench of protein fluorescence. The binding of NADH to the enzyme was pH-dependent. At pH8.1 a biphasic profile was obtained on titrating the enzyme with NADH, whereas at pH8.8 the titration profile was hyperbolic. UDP-xylose, and to a lesser extent UDP-glucuronic acid, lowered the apparent affinity of the enzyme for NADH.  相似文献   

5.
An NAD(+)-dependent CDP-D-glucose oxidoreductase which catalyzes the first step of the biosynthesis of CDP-ascarylose (CDP-3,6-dideoxy-L-arabino-hexose), converting CDP-D-glucose to CDP-4-keto-6-deoxy-D-glucose, was isolated from Yersinia pseudotuberculosis. A protocol consisting of DEAE-cellulose, Matrex Blue-A, hydroxylapatite, DEAE-Sephadex, Sephadex G-100, and NAD(+)-agarose column chromatography was used to purify this enzyme 6000-fold to homogeneity. This enzyme consists of two identical subunits, each with a molecular weight of 42,500. Using CDP-D-glucose as the substrate, the Km and Vmax of this catalysis were determined to be 222 microM and 8.3 mumols mg-1 min-1, respectively. Unlike most other oxidoreductases of its class which have a tightly bound NAD+, this highly purified CDP-D-glucose oxidoreductase showed an absolute requirement of NAD+ for its activity. Using chemically synthesized (6S)- and (6R)-CDP-D-[4-2H,6-3H]glucose as substrates, a stereochemical analysis showed this enzymatic reaction involves an intramolecular hydrogen migration from C-4 to C-6, and the displacement of C-6 hydroxyl group by the C-4 hydrogen occurs with inversion. Thus, despite the low cofactor affinity, this enzyme undergoes a mechanism consistent with that followed by other members of its type. Such a mechanistic and stereochemical convergency found for all sugar oxidoreductases so far characterized suggests the presence of a common progenitor of this class of enzyme.  相似文献   

6.
Using ammonium sulfate precipitation, gel filtration, and affinity chromatography, inosine monophosphate (IMP) oxidoreductase (EC 1.2.1.14) was isolated from the soluble proteins of the plant cell fraction of nitrogen-fixing nodules of cowpea (Vigna unguiculata L. Walp). The enzyme, purified more than 140-fold with a yield of 11%, was stabilized with glycerol and required a sulfydryl-reducing agent for maximum activity. Gel filtration indicated a molecular weight of 200,000, and sodium dodecyl sulfate-gel electrophoresis a single subunit of 50,000 Da. The final specific activity ranged from 1.1 to 1.5 mumol min-1 mg protein-1. The enzyme had an alkaline pH optimum and showed a high affinity for IMP (Km = 9.1 X 10(-6) M at pH 8.8 and NAD levels above 0.25 mM) and NAD (Km = 18-35 X 10(-6) M at pH 8.8). NAD was the preferred coenzyme, with NADP reduction less than 10% of that with NAD, while molecular oxygen did not serve as an electron acceptor. Intermediates of ureide metabolism (allantoin, allantoic acid, uric acid, inosine, xanthosine, and XMP) did not affect the enzyme, while AMP, GMP, and NADH were inhibitors. GMP inhibition was competitive with a Ki = 60 X 10(-6) M. The purified enzyme was activated by K+ (Km = 1.6 X 10(-3) M) but not by NH+4. The K+ activation was competitively inhibited by Mg2+. The significance of the properties of IMP oxidoreductase for regulation of ureide biosynthesis in legume root nodules is discussed.  相似文献   

7.
1. A ;malic' enzyme [l-malate-NAD oxidoreductase (decarboxylating), EC 1.1.1.39] has been isolated from cauliflower bud mitochondria and partially purified. 2. The enzyme is specific for l-malate and has an absolute requirement for either Mn(2+), Co(2+) or Mg(2+). 3. The enzyme shows activity with both NAD(+) and NADP(+), but NAD(+) is the preferred cofactor. 4. No appreciable oxaloacetate decarboxylase activity is present in the enzyme preparations even at low pH values. 5. The enzyme is inhibited by NADH and by oxaloacetate and stimulated by SO(4) (2-) and by low concentrations of CoA. 6. The regulatory properties of the enzyme support the proposed role of the enzyme in the utilization of tricarboxylic acid-cycle acids for energy production when glycolysis is suppressed.  相似文献   

8.
When oxidized to cysteic acid by performic acid or converted to carboxymethylcysteine by alkylation of the reduced enzyme with iodoacetate, a total of six half-cystine residues/subunit are found in L-threonine dehydrogenase (L-threonine: NAD+ oxidoreductase, EC 1.1.1.103; L-threonine + NAD(+)----2-amino-3-oxobutyrate + NADH) from Escherichia coli K-12. Of this total, two exist in disulfide linkage, whereas four are titratable under denaturing conditions by dithiodipyridine, 5,5'-dithiobis(2-nitrobenzoic acid), or p-mercuribenzoate. The kinetics of enzyme inactivation and of modification by the latter two reagents indicate that threonine dehydrogenase has no free thiols that selectively react with bulky compounds. While incubation of the enzyme with a large excess of iodoacetamide causes less than 10% loss of activity, the native dehydrogenase is uniquely reactive with and completely inactivated by iodoacetate. The rate of carboxymethylation by iodoacetate of one -SH group/subunit is identical with the rate of inactivation and the carboxymethylated enzyme is no longer able to bind Mn2+. NADH (0.5 mM) provides 40% protection against this inactivation; 60 to 70% protection is seen in the presence of saturating levels of NADH plus L-threonine. Such results coupled with an analysis of the kinetics of inactivation caused by iodoacetate are interpreted as indicating the inhibitor first forms a reversible complex with a positively charged moiety in or near the microenvironment of a reactive -SH group in the enzyme before irreversible alkylation occurs. Specific alkylation of one -SH group/enzyme subunit apparently causes protein conformational changes that entail a loss of catalytic activity and the ability to bind Mn2+.  相似文献   

9.
Moorella thermoacetica ferments glucose to three acetic acids. In the oxidative part of the fermentation, the hexose is converted to 2 acetic acids and 2 CO(2) molecules with the formation of 2 NADH and 2 reduced ferredoxin (Fd(red)(2-)) molecules. In the reductive part, 2 CO(2) molecules are reduced to acetic acid, consuming the 8 reducing equivalents generated in the oxidative part. An open question is how the two parts are electronically connected, since two of the four oxidoreductases involved in acetogenesis from CO(2) are NADP specific rather than NAD specific. We report here that the 2 NADPH molecules required for CO(2) reduction to acetic acid are generated by the reduction of 2 NADP(+) molecules with 1 NADH and 1 Fd(red)(2-) catalyzed by the electron-bifurcating NADH-dependent reduced ferredoxin:NADP(+) oxidoreductase (NfnAB). The cytoplasmic iron-sulfur flavoprotein was heterologously produced in Escherichia coli, purified, and characterized. The purified enzyme was composed of 30-kDa (NfnA) and 50-kDa (NfnB) subunits in a 1-to-1 stoichiometry. NfnA harbors a [2Fe2S] cluster and flavin adenine dinucleotide (FAD), and NfnB harbors two [4Fe4S] clusters and FAD. M. thermoacetica contains a second electron-bifurcating enzyme. Cell extracts catalyzed the coupled reduction of NAD(+) and Fd with 2 H(2) molecules. The specific activity of this cytoplasmic enzyme was 3-fold higher in H(2)-CO(2)-grown cells than in glucose-grown cells. The function of this electron-bifurcating hydrogenase is not yet clear, since H(2)-CO(2)-grown cells additionally contain high specific activities of an NADP(+)-dependent hydrogenase that catalyzes the reduction of NADP(+) with H(2). This activity is hardly detectable in glucose-grown cells.  相似文献   

10.
NAD(+)-dependent isocitrate dehydrogenase from Saccharomyces cerevisiae is composed of two nonidentical subunits, designated IDH1 and IDH2. The gene encoding IDH2 was previously cloned and sequenced (Cupp, J.R., and McAlister-Henn, L. (1991) J. Biol. Chem. 266, 22199-22205), and in this paper we describe the isolation of a yeast genomic clone containing the IDH1 gene. A fragment of the IDH1 gene was amplified by the polymerase chain reaction method utilizing degenerate oligonucleotides based on tryptic peptide sequences of the purified subunit; this fragment was used to isolate a full length IDH1 clone. The nucleotide sequence of the IDH1 coding region was determined and encodes a 360-residue polypeptide including an 11-residue mitochondrial targeting presequence. Amino acid sequence comparison between IDH1 and IDH2 reveals a 42% sequence identity, and both IDH1 and IDH2 show approximately 32% identity to Escherichia coli NAD(P)(+)-dependent isocitrate dehydrogenase. To examine the function of the IDH1 subunit and to determine the metabolic role of NAD(+)-dependent isocitrate dehydrogenase the IDH1 gene was disrupted in a wild type haploid yeast strain and in a haploid strain lacking IDH2. The IDH1 disruption strains expressed no detectable IDH1 as determined by Western blot analysis, and these strains were found to lack NAD(+)-dependent isocitrate dehydrogenase activity indicating that IDH1 is essential for a functional enzyme. Over-expression of IDH1 in a strain containing IDH2 restored wild type activity but did not result in increased levels of activity, suggesting that both IDH1 and IDH2 are required for a functional enzyme. Growth phenotype analysis of the IDH1 disruption strains revealed that they grew at a reduced rate on the nonfermentable carbon sources examined (glycerol, lactate, and acetate), consistent with NAD(+)-dependent isocitrate dehydrogenase performing a critical role in oxidative function of the citric acid cycle. In addition, the IDH1 disruption strains grew at wild type rates in the absence of glutamate, indicating that these strains are not glutamate auxotrophs.  相似文献   

11.
Glycerol-3-phosphate oxidoreductase (sn-glycerol 3-phosphate: NAD+ 2-oxidoreductase, EC 1.1.1.8) from human placenta has been purified by chromatography on 2,4,6-trinitrobenzenehexamethylenediamine-Sepharose, DEAE-Sephadex A-50 and 5'-AMP-Sepharose 4B approximately 15800-fold with an overall yield of about 19%. The final purified material displayed a specific activity of about 88 mumol NADH min-1 mg protein-1 and a single protein band on polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulphate. The native molecular mass, determined by Ultrogel AcA 44 filtration, was 62000 +/- 2000 whereas the subunit molecular mass, established on polyacrylamide gel in the presence of 0.1% sodium dodecyl sulphate, was 38000 +/- 500. The isoelectric point of the enzyme protein, determined by column isoelectric focusing, was found to be 5.29 +/- 0.09. The pH optimum of the placental enzyme was in the range 7.4-8.1 for dihydroxyacetone phosphate reduction and 8.7-9.2 for sn-glycerol 3-phosphate oxidation. The apparent Michaelis constants (Km) for dihydroxyacetone phosphate, NADH, sn-glycerol 3-phosphate and NAD+ were 26 microM, 5 microM, 143 microM and 36 microM respectively. The activity ratio of cytoplasmic glycerol-3-phosphate oxidoreductase to mitochondrial glycerol-3-phosphate dehydrogenase in human placental tissue was 1:2. The consumption of oxygen by human placental mitochondria incubated with the purified glycerol-3-phosphate oxidoreductase, NADH and dihydroxyacetone phosphate was similar to that observed in the presence of sn-glycerol 3-phosphate. The possible physiological role of glycerol-3-phosphate oxidoreductase in placental metabolism is discussed.  相似文献   

12.
1. An enzyme system that catalyses the alpha-oxidation of fatty acids to shorter chain products is present in acetone powders of cucumber fruits. 2. In the absence of NAD+, the predominant product from palmitic acid is pentadecanal. Addition of NAD+ gives rise to a homologous series of n-alkanals, the concentrations of which are in the same order as that reported in the volatile products formed on homogenization of cucumbers, i.e. C15 greater than C14 greater than C13 greater than C12. 3. Pentadecan-1-ol is also formed from palmitic acid in the absence of added NAD+; C15, C14 and C13 n-alkanols are produced in the presence of NAD+. 4. The substrate specificity for saturated fatty acids is in the order C12 less than C14 greater than C16 greater than C18. Unsaturated C18 acids are oxidized more readily than stearic acid. 5. The alpha-oxidation system is inhibited by dithiothreitol, cysteine, imidazole and certain metal ligands (CN-, N3-, diphenylthiocarbazone) but not by EDTA. 6. Differences between the alpha-oxidation system in cucumber and those previously reported in other plants are discussed.  相似文献   

13.
The NADP(+)-preferring glucose dehydrogenase from thermoacidophilic archaeon Thermoplasma acidophilum has been characterized, and its crystal structure has been determined (Structure, 2:385-393, 1994). Its sequence and structure are not homologous to bacterial NAD(P)(+)-dependent glucose dehydrogenases, and its molecular weight is also quite defferent. On the other hand, three functionally unknown genes with homologies to bacterial NAD(P)(+)-dependent glucose dehydrogenases have been sequenced as part of the T. acidophilum genome project (gene names: Ta0191, Ta0747, and Ta0754 respectively). We expressed two genes of three, Ta0191 and Ta0754, in Escherichia coli, and purified the gene products to homogeneity. Dehydrogenase activities were thereby detected from the purified proteins. The Ta0754 gene product exhibited aldohexose dehydrogenase activity, and the Ta0191 gene product exhibited weak 2-deoxyglucose dehydrogenase activity. No aldohexose dehydrogenase gene has been isolated, while the enzyme was reported in 1968. This is the first report of the gene and primary structure. The purified Ta0754 gene product, designated AldT, was characterized. The enzyme AldT effectively catalyzed the oxidation of various aldohexoses, especially D-mannose. Lower activities on D-2-deoxyglucose, D-xylose, D-glucose, and D-fucose were detected although no activities were shown on other aldohexoses or additional sugars. As a cofactor, NAD(+) was much more suitable for the activity than NADP(+). The NAD(+)-preferring dehydrogenase most effectively reacting to D-mannose is for the first time. AldT was most active at pH 10 and above 70 degrees C, and completely stable up to 60 degrees C after incubation for 15 min. Other enzymatic properties were also investigated.  相似文献   

14.
The essential gene efg, which complements ammonia-dependent growth (adgA) mutations in Rhodobacter capsulatus and is located at 38.1 min on the Escherichia coli chromosome, was found to code for NH3-dependent NAD synthetase. Crude extracts from a strain which overproduces the efg gene product contained up to 400 times more activity than crude extracts from the control strain, and the purified Efg protein possessed-NH3-dependent NAD synthetase activity. Glutamine-dependent NAD synthetase activity was found in crude extracts of E. coli but not in the purified enzyme, suggesting that it may be catalyzed by an additional subunit. An R. capsulatus strain carrying an adgA mutation was found to be deficient in NAD synthetase activity, and activity was restored by complementation with the E. coli gene. In accordance with the nomenclature proposed for Salmonella typhimurium (K. T. Hughes, B. M. Olivera, and J. R. Roth, J. Bacteriol. 170:2113-2120, 1988), the efg and adgA genes should now be designated nadE.  相似文献   

15.
When 20-hydroxyleukotriene B4 (20-OH-LTB4) is incubated at pH 10.5 in the presence of NAD+ with an alcohol dehydrogenase isolated from human neutrophils, a polar product is formed as detected on reverse-phase high-performance liquid chromatography (RP-HPLC). The product is identified as 20-oxo-LTB4 (20-CHO-LTB4) on the basis of its co-elution with the authentic compound on HPLC, ultraviolet spectrometry and gas chromatography-mass spectrometry. The 20-CHO-LTB4-forming activity requires NAD+, but NADP+ scarcely replaces NAD+. The apparent Km for 20-OH-LTB4 is 83 microM and the Vmax is 2.04 mumol/min per mg of protein. The activity is inhibited by omega-hydroxy fatty acids such as 12-hydroxylauric acid, 16-hydroxypalmitic acid and 12(S), 20-dihydroxyeicosatetraenoic acid, but not by 4-methylpyrazole. At pH 7.0 with NADH, the purified dehydrogenase catalyzes the reverse reaction, the reduction of 20-CHO-LTB4 to 20-OH-LTB4.  相似文献   

16.
3-Hydroxyisobutyrate dehydrogenase (3-hydroxy-2-methyl propanoate: NAD+ oxidoreductase, EC 1.1.1.31) was purified 1800-fold from rabbit liver by detergent extraction, differential solubility in polyethylene glycol and (NH4)2SO4, and column chromatography on DEAE-Sephacel, phenyl-Sepharose, CM(carboxymethyl)-Sepharose, Affi-Gel Blue, and Ultrogel AcA-34. The enzyme had a native Mr of 74,000 and appeared to be a homodimer with subunit Mr = 34,000. The enzyme was specific for NAD+. It oxidized both S-3-hydroxyisobutyrate and R-3-hydroxyisobutyrate, but the kcat/Km was approximately 350-fold higher for the S-isomer. Steady state kinetic analysis indicates an ordered Bi Bi reaction mechanism with NAD+ binding before 3-hydroxyisobutyrate. The enzyme catalyzed oxidation of S-3-hydroxyisobutyrate between pH 7.0 and 11.5 with optimal activity between pH 9.0 and 11.0. The enzyme apparently does not have a metal ion requirement. Essential sulfhydryl groups may be present at both the 3-hydroxyisobutyrate and NAD+ binding sites since inhibition by sulfhydryl-binding agents was differentially blocked by each substrate. The enzyme is highly sensitive to product inhibition by NADH which may play an important physiological role in regulating the complete oxidation of valine beyond the formation of 3-hydroxyisobutyrate.  相似文献   

17.
Heterotetrameric (alphabetagammadelta) sarcosine oxidase from Corynebacterium sp. P-1 (cTSOX) contains noncovalently bound FAD and NAD(+) and covalently bound FMN, attached to beta(His173). The beta(His173Asn) mutant is expressed as a catalytically inactive, labile heterotetramer. The beta and delta subunits are lost during mutant enzyme purification, which yields a stable alphagamma complex. Addition of stabilizing agents prevents loss of the delta but not the beta subunit. The covalent flavin link is clearly a critical structural element and essential for TSOX activity or preventing FMN loss. The alpha subunit was expressed by itself and purified by affinity chromatography. The alpha and beta subunits each contain an NH(2)-terminal ADP-binding motif that could serve as part of the binding site for NAD(+) or FAD. The alpha subunit and the alphagamma complex were each found to contain 1 mol of NAD(+) but no FAD. Since NAD(+) binds to alpha, FAD probably binds to beta. The latter could not be directly demonstrated since it was not possible to express beta by itself. However, FAD in TSOX from Pseudomonas maltophilia (pTSOX) exhibits properties similar to those observed for the covalently bound FAD in monomeric sarcosine oxidase and N-methyltryptophan oxidase, enzymes that exhibit sequence homology with beta. A highly conserved glycine in the ADP-binding motif of the alpha(Gly139) or beta(Gly30) subunit was mutated in an attempt to generate NAD(+)- or FAD-free cTSOX, respectively. The alpha(Gly139Ala) mutant is expressed only at low temperature (t(optimum) = 15 degrees C), but the purified enzyme exhibited properties indistinguishable from the wild-type enzyme. The much larger barrier to NAD(+) binding in the case of the alpha(Gly139Val) mutant could not be overcome even by growth at 3 degrees C, suggesting that NAD(+) binding is required for TSOX expression. The beta(Gly30Ala) mutant exhibited subunit expression levels similar to those of the wild-type enzyme, but the mutation blocked subunit assembly and covalent attachment of FMN, suggesting that both processes require a conformational change in beta that is induced upon FAD binding. About half of the covalent FMN in recombinant preparations of cTSOX or pTSOX is present as a reversible covalent 4a-adduct with a cysteine residue. Adduct formation is not prevented by mutating any of the three cysteine residues in the beta subunit of cTSOX to Ser or Ala. Since FMN is attached via its 8-methyl group to the beta subunit, the FMN ring must be located at the interface between beta and another subunit that contains the reactive cysteine residue.  相似文献   

18.
Summary Plasma membranes isolated from onion roots by twophase partition contain at least two different NAD(P)H-dehydrogenases. A 27 kDa electron transport protein oxidises both NADH and NADPH and exhibits maximal activity with quinones as electron acceptors. A distinct 31 kDa dehydrogenase is specific for NADH as donor and shows maximal activity with ferricyanide. This novel enzyme is responsible for most NADH-ferricyanide oxidoreductase activity of solubilized onion root plasma membranes and exhibits properties different to other purified NAD(P)H-dehydrogenases.Abbreviations DES diethylstilbestrol - FeCN potassium ferricyanide - NBT nitroblue tetrazolium - PHMB p-hydroxymercuribenzoate - PMSF phenylmethylsulfonylfluoride - PTA phosphotungstic acid - SHAM salicylhydroxamic acid - TTFA thenoyltrifluoroacetone  相似文献   

19.
Aminopropionaldehyde dehydrogenase was purified to apparent homogeneity from 1,3-diaminopropane-grown cells of Arthrobacter sp. TMP-1. The native molecular mass and the subunit molecular mass of the enzyme were approximately 20,5000 and 52,000, respectively, suggesting that the enzyme is a tetramer of identical subunits. The apparent Michaelis constant (K(m)) for 1,3-diaminopropane was approximately 3 microM. The enzyme equally used both NAD(+) and NADP(+) as coenzymes. The apparent K(m) values for NAD(+) and NADP(+) were 255 microM and 108 microM, respectively. The maximum reaction rates (V(max)) for NAD(+) and NADP(+) were 102 and 83.3 micromol min(-1) mg(-1), respectively. Some tested aliphatic aldehydes and aromatic aldehydes were inert as substrates. The optimum pH was 8.0-8.5. The enzyme was sensitive to sulfhydryl group-modifying reagents.  相似文献   

20.
The course of the reaction sequence hypoxanthine leads to xanthine leads to uric acid, catalysed by the NAD+-dependent activity of xanthine oxidoreductase, was investigated under conditions either of immediate oxidation of the NADH formed or of NADH accumulation. The enzymic preparation was obtained from rat liver, and purified 75-fold (as compared with the 25000 g supernatant) on a 5'-AMP-Sepharose 4B column; in this preparation the NAD+-dependent activity accounted for 100% of total xanthine oxidoreductase activity. A spectrophotometric method was developed for continuous measurements of changes in the concentrations of the three purines involved. The time course as well as the effects of the concentrations of enzyme and of hypoxanthine were examined. NADH produced by the enzyme lowered its activity by 50%, resulting in xanthine accumulation and in decreases of uric acid formation and of hypoxanthine utilization. The inhibition of the Xanthine oxidoreductase NAD+-dependent activity by NADH is discussed as a possible factor in the regulation of IMP biosynthesis by the 'de novo' pathway or (from unchanged hypoxanthine) by ther salvage pathway.  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