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.
Soluble formate dehydrogenase from Methanobacterium formicicum was purified 71-fold with a yield of 35%. Purification was performed anaerobically in the presence of 10 mM sodium azide which stabilized the enzyme. The purified enzyme reduced, with formate, 50 mumol of methyl viologen per min per mg of protein and 8.2 mumol of coenzyme F420 per min per mg of protein. The apparent Km for 7,8-didemethyl-8-hydroxy-5-deazariboflavin, a hydrolytic derivative of coenzyme F420, was 10-fold greater (63 microM) than for coenzyme F420 (6 microM). The purified enzyme also reduced flavin mononucleotide (Km = 13 microM) and flavin adenine dinucleotide (Km = 25 microM) with formate, but did not reduce NAD+ or NADP+. The reduction of NADP+ with formate required formate dehydrogenase, coenzyme F420, and coenzyme F420:NADP+ oxidoreductase. The formate dehydrogenase had an optimal pH of 7.9 when assayed with the physiological electron acceptor coenzyme F420. The optimal reaction rate occurred at 55 degrees C. The molecular weight was 288,000 as determined by gel filtration. The purified formate dehydrogenase was strongly inhibited by cyanide (Ki = 6 microM), azide (Ki = 39 microM), alpha,alpha-dipyridyl, and 1,10-phenanthroline. Denaturation of the purified formate dehydrogenase with sodium dodecyl sulfate under aerobic conditions revealed a fluorescent compound. Maximal excitation occurred at 385 nm, with minor peaks at 277 and 302 nm. Maximal fluorescence emission occurred at 455 nm.  相似文献   

2.
Organoautotrophic growth of Alcaligenes eutrophus on formate was dependent on the presence of molybdate in the medium. Supplementation of the medium with tungstate lead to growth cessation. Corresponding effects of these anions were observed for the activity of the soluble, NAD(+)-linked formate dehydrogenase (S-FDH; EC 1.2.1.2) of the organism. Lack of molybdate or presence of tungstate resulted in an almost complete loss of S-FDH activity. S-FDH was purified to near homogeneity in the presence of nitrate as a stabilizing agent. The native enzyme exhibited an M(r) of 197,000 and a heterotetrameric quaternary structure with nonidentical subunits of M(r) 110,000 (alpha), 57,000 (beta), 19,400 (gamma), and 11,600 (delta). It contained 0.64 g-atom of molybdenum, 25 g-atom of nonheme iron, 20 g-atom of acid-labile sulfur, and 0.9 mol of flavin mononucleotide per mol. The fluorescence spectrum of iodine-oxidized S-FDH was nearly identical to the form A spectrum of milk xanthine oxidase, proving the presence of a pterin cofactor. The molybdenum-complexing cofactor was identified as molybdopterin guanine dinucleotide in an amount of 0.71 mol/mol of S-FDH. Apparent Km values of 3.3 mM for formate and 0.09 mM for NAD+ were determined. The enzyme coupled the oxidation of formate to a number of artificial electron acceptors and was strongly inactivated by formate in the absence of NAD+. It was inhibited by cyanide, azide, nitrate, and Hg2+ ions. Thus, the enzyme belongs to a new group of complex molybdo-flavo Fe-S FDH that so far has been detected in only one other aerobic bacterium.  相似文献   

3.
A cyanide-degrading enzyme from Bacillus pumilus C1 has been purified and characterized. This enzyme consisted of three polypeptides of 45.6, 44.6, and 41.2 kDa; the molecular mass by gel filtration was 417 kDa. Electron microscopy revealed a multimeric, rod-shaped protein approximately 9 by 50 nm. Cyanide was rapidly degraded to formate and ammonia. Enzyme activity was optimal at 37 degrees C and pH 7.8 to 8.0. Activity was enhanced by Sc3+, Cr3+, Fe3+, and Tb3+; enhancement was independent of metal ion concentration at concentrations above 5 microM. Reversible enhancement of enzymatic activity by azide was maximal at 4.5 mM azide and increased with time. No activity was recorded with the cyanide substrate analogs CNO-, SCN-, CH3CN, and N3- and the possible degradation intermediate HCONH2. Kinetic studies indicated a Km of 2.56 +/- 0.48 mM for cyanide and a Vmax of 88.03 +/- 4.67 mmol of cyanide per min/mg/liter. The Km increased approximately twofold in the presence of 10 microM Cr3+ to 5.28 +/- 0.38 mM for cyanide, and the Vmax increased to 197.11 +/- 8.51 mmol of cyanide per min/mg/liter. We propose naming this enzyme cyanide dihydratase.  相似文献   

4.
Two previous kinetic studies on the Arabidopsis thaliana leaf NAD-dependent formate dehydrogenase (EC 1.2.1.2) have demonstrated two very different sets of Km values for the formate and NAD+ substrates. We examined the kinetics of the enzyme partially purified from a leaf extract by gel-filtration desalting and chromatography on DEAE-cellulose, as well as by isolation of a mitochondria-enriched fraction obtained by differential centrifugation. Both of these methods produce a formate dehydrogenase enzyme with the higher Km values of approximately 10 mmol/L formate and 75 mumol/L NAD+. The kinetic properties of the Arabidopsis formate dehydrogenase expressed to high levels in transgenic tobacco plants were also those of the high Km form. The high Km form of the enzyme converted to a low Km form by heating for 5 minutes at 60 degrees C. An Arrhenius plot of the activity during the heating process was linear, indicating that the heating did not cause alterations in either the active site or the thermal dependence of the catalytic reaction. We conclude that the native form of the formate dehydrogenase probably resembles the form with the higher Km values. Heating seemingly converts this native enzyme to the molten globule state and cooling results in formation of a non-native structure with altered kinetic properties.  相似文献   

5.
Succinate semialdehyde dehydrogenase (SSADH) has been purified from potato tubers with 39% yield, 832-fold purification, and a specific activity of 6.5 units/mg protein. The final preparation was homogeneous as judged from native and sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Gel filtration on Sepharose 6B gave a relative molecular mass (Mr) of 145,000 for the native enzyme. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis gave a single polypeptide band of Mr 35,000. Thus the enzyme appears to be a tetramer of identical subunits. Chromatofocusing of the enzyme gave a pI of 8.7. The enzyme was maximally active at pH 9.0 in 100 mM sodium pyrophosphate buffer. In 100 mM Tris-HCl buffer, pH 9.0, the enzyme gave only 20% of the activity found in pyrophosphate buffer and had a shorter linear rate. The enzyme was specific for succinate semialdehyde (SSA) as substrate and could not utilize acetaldehyde, glyceraldehyde 3-phosphate, malonaldehyde, lactate, or ethanol as substrates. The enzyme was also specific for NAD+ as cofactor and NADP+ and 3-acetylpyridine adenine dinucleotide could not serve as cofactors. Potato SSADH had a Km of 4.6 microM for SSA when assayed in pyrophosphate buffer and was inhibited by that substrate at concentrations greater than 120 microM. The Km for NAD+ was found to be 31 microM. The enzyme required exogenous addition of a thiol compound for maximal activity and was inhibited by the thiol-directed reagents p-hydroxymercuribenzoate, dithionitrobenzoate, and N-ethyl-maleimide, by heavy metal ions Hg2+, Cu2+, Cd2+, and Zn2+, and by arsenite. These results indicate a requirement of a SH group for catalytic activity.  相似文献   

6.
A ferredoxin-NAD+ oxidoreductase (EC 1.18.1.3) has been isolated from extracts of the obligate methanotroph Methylosinus trichosporium OB3b. This enzyme was shown to couple electron flow from formate dehydrogenase (NAD+ requiring) to ferredoxin. Ferredoxin-NAD+ reductase was purified to homogeneity by conventional chromatography techniques and was shown to be a flavoprotein with a molecular weight of 36,000 +/- 1,000. This ferredoxin reductase was specific for NADH (Km, 125 microM) and coupled electron flow to the native ferredoxin and to ferredoxins from spinach, Clostridium pasteurianum, and Rhodospirillum rubrum (ferredoxin II). M. trichosporium ferredoxin saturated the ferredoxin-NAD+ reductase at a concentration 2 orders of magnitude lower (3 nM) than did spinach ferredoxin (0.4 microM). Ferredoxin-NAD+ reductase also had transhydrogenase activity which transferred electrons and protons from NADH to thionicotinamide adenine dinucleotide phosphate (Km, 9 microM) and from NADPH to 3-acetylpyridine adenine dinucleotide (Km, 16 microM). Reconstitution of a soluble electron transport pathway that coupled formate oxidation to ferredoxin reduction required formate dehydrogenase, NAD+, and ferredoxin-NAD+ reductase.  相似文献   

7.
The formate dehydrogenase from the yeast Pichia pastoris IFP 206 was purified to homogeneity. The protein showed a molecular weight of 68,000 daltons and was composed of two identical subunits. Its amino acid composition was similar to those of other formate dehydrogenases and was characterized by a high content of acidic residues. The N-terminal end of the molecule was probably blocked.

The enzyme activity was NAD+ dependent (NADP+ could not replace NAD+). Its optimum temperature was 47°C and the activation energy 10.8 kcal/mol. The enzyme was active from pH 3.5 to 10.5 with a maximum at pH 7.5. The Michaelis constant for NAD+ and formate were respectively 0.27 and 15mM. The purified enzyme had no S-formylglutathione hydrolase activity, strongly suggesting that the true substrate was formate. NADH, cyanide and azide were strong inhibitors of the enzyme.  相似文献   

8.
Benzaldehyde dehydrogenase I was purified from Acinetobacter calcoaceticus by DEAE-Sephacel, phenyl-Sepharose and f.p.l.c. gel-filtration chromatography. The enzyme was homogeneous and completely free from the isofunctional enzyme benzaldehyde dehydrogenase II, as judged by denaturing and non-denaturing polyacrylamide-gel electrophoresis. The subunit Mr value was 56,000 (determined by SDS/polyacrylamide-gel electrophoresis). Estimations of the native Mr value by gel-filtration chromatography gave values of 141,000 with a f.p.l.c. Superose 6 column, but 219,000 with Sephacryl S300. Chemical cross-linking of the enzyme subunits indicated that the enzyme is tetrameric. Benzaldehyde dehydrogenase I was activated more than 100-fold by K+, Rb+ and NH4+, and the apparent Km for K+ was 11.2 mM. The pH optimum in the presence of K+ was 9.5 and the pI of the enzyme was 5.55. The apparent Km values for benzaldehyde and NAD+ were 0.69 microM and 96 microM respectively, and the maximum velocity was approx. 110 mumol/min per mg of protein. Various substituted benzaldehydes were oxidized at significant rates, and NADP+ was also used as cofactor, although much less effectively than NAD+. Benzaldehyde dehydrogenase I had an NAD+-activated esterase activity with 4-nitrophenol acetate as substrate, and the dehydrogenase activity was inhibited by a range of thiol-blocking reagents. The absorption spectrum indicated that there was no bound cofactor or prosthetic group. Some of the properties of the enzyme are compared with those of other aldehyde dehydrogenases, specifically the very similar isofunctional enzyme benzaldehyde dehydrogenase II from the same organism.  相似文献   

9.
Modification of two SH-groups in the molecule of formate dehydrogenase by dithiobisnitrobenzoate or to dacetamide results in the enzyme inactivation. Coenzymes, but not the substrate, protect the enzyme against the inactivation. NAD in the presence of potassium azide completely preserves the enzyme activity. Two SH-groups per enzyme molecule are protected from modification. The Km values for partially inactivated formate dehydrogenase remain constant for both substrates. The enzyme with modified SH-groups does not bind conezymes. The pH-dependence of the inactivation rate reveals the ionizable group with pK 9.6 (25 degrees C). The involvement of essential SH-groups in coenzyme binding is discussed.  相似文献   

10.
The purified respiratory chain NADH dehydrogenase of Escherichia coli oxidizes NADH with either dichlorophenolindophenol (DCIP). ferricyanide, or menadione as electron acceptors, with values for NADH are similar with the three electron acceptors (approximately 50 muM). The purified enzyme contains no flavin and has an absolute requirement for FAD, with Km values around 4 muM. The pH optimum of the enzyme appears to be between 6.5 and 7; the optimum is difficult to establish because of nonenzymatic reduction of DCIP at the lower pH values. Potassium cyanide stimulates the DCIP reductase activity about 2-fold, but has no effect on ferricyanide reductase. The enzyme exhibits hyperbolic kinetics with respect to NADH concentration in both the ferricyanide and DCIP reductase assays, but cooperatively is seen in the menadione reductase reaction. NAD+ is an effective competitive inhibitor of the reaction (Ki congruent to 20 muM); in the presence of NAD+, the NADH saturation curve becomes cooperative, even in the DCIP reductase assay. Many adenine containing nucleotides are competitive inhibitors of the enzyme. The apparent Ki values for these nucleotides as inhibitors of the purified enzyme, the membrane-bound NADH dehydrogenase, and the NADH oxidase are equivalent. An examination of inhibitory effects of a series of adenine nucleotides suggests that the inhibitors act as analogues of NAD+, which is the true physiological inhibitor. The results suggest that the enzyme in situ is always partially inhibited by the levels of NAD- in the E coli cell, and thus behaves in a cooperative fashion to changes in the NAD+/NADH ratio. An antibody has been elicited against the purified NADH dehydrogenase. Immunodiffusion and crossed immunoelectrophoresis show that the antibody is directed principally against the NADH dehydrogenase, with some activity against minor contaminants in the purified preparation. The antibody inhibits NADH dehydrogenase activity 50% at saturating levels. When this antibody preparation is used to examine solubilized membrane preparations, two major immunoprecipitates are found. A parallel inhibition of the membrane-bound NADH dehydrogenase and NADH oxidase activities is seen, supporting the hypothesis that the purified enzyme is indeed a component of the respiratory chain-dependent NADH oxidase pathway.  相似文献   

11.
Mycobacterium vaccae 10 growing in methanol medium synthesizes two inducible alternative NAD(+)-dependent formate dehydrogenases (FDH). In the presence of molybdenum, the dominating form of the enzyme is FDHI with Mr 440 kDa and Km 0.32 mM for sodium formate. FDHI reduced ferricyanide as well as NAD+, and it was reversibly inactivated by formate. NAD+ stabilized FDHI against this inactivation. Under conditions of artificial molybdenum deficiency (tungsten in the medium), the second enzyme (FDHII) appeared with Mr about 93 kDa and Km 8.3 mM for sodium formate, and no FDHI activity was detected. FDHII did not reduce ferricyanide and was not inactivated by formate. The activity of FDHI was restored in tungsten-grown cells by pulse addition of molybdenum under conditions of blocked protein synthesis, suggesting the pre-existence of inactive apo-FDHI.  相似文献   

12.
The catalytic activity, expressed as Km and Vmax values, of 16 enzymes of practical interest with the macromolecular coenzymes poly(ethylene glycol)-N6-(2-aminoethyl)-NAD+ and poly(ethylene glycol)-N6-(2-aminoethyl)-NADP+ and their low molecular weight precursors N6-(2-aminoethyl)-NAD+ and N6-(2-aminoethyl)-NADP+, was investigated. The enzymes examined are of direct interest for organic synthesis (i.e. alcohol dehydrogenase from yeast, horse liver, or Thermoanaerobium brockii, lactic dehydrogenase, and several hydroxysteroid dehydrogenases) or are used for the regeneration of NAD+, NADP+, NADH, or NADPH (i.e. glutamate dehydrogenase from liver or Proteus, formate dehydrogenase, glucose dehydrogenase, and malic enzyme). The cycling efficiency of poly(ethylene glycol)-N6-(2-aminoethyl)-NADP+ was examined with coupled-enzymes or coupled-substrates systems. Poly(ethylene glycol)-N6-(2-aminoethyl)-NAD+ and, even more so, poly(ethylene glycol)-N6-(2-aminoethyl)-NADP+ were excellent coenzymes with several dehydrogenases. In addition, the coenzymatic properties of N6-(3-sulfonatopropyl)-NAD+, an NAD+ derivative carrying a strong anionic group, were compared with those of the newly synthesized N6-(2-hydroxy-3-trimethylammonium propyl)-NAD+, an NAD+ derivative carrying a strong cationic group. It was expected that the presence of the sulfonic or quaternary ammonium group would enhance the residence time of the coenzyme inside continuous-flow reactors if membranes with anionic or cationic groups, respectively, were used.  相似文献   

13.
NAD+-dependent formate dehydrogenase was screened in various bacterial strains. Facultative methanol-utilizing bacteria isolated from soil samples, acclimated to a medium containing methanol and formate at pH 9.5, were classified as members of the genus Moraxella. From a crude extract of Moraxella sp. strain C-1, formate dehydrogenase was purified to homogeneity, as judged by disc gel electrophoresis. The enzyme has an isoelectric point of 3.9 and a molecular weight of approximately 98,000. The enzyme is composed of two identical subunits with molecular weights of about 48,000. The apparent Km values for sodium formate and NAD+ were calculated to be 13 mM and 0.068 mM, respectively.  相似文献   

14.
Pork liver 2-keto-3-deoxy-L-fuconate:NAD+ oxidoreductase has been shown to convert 2-keto-3-deoxy-L-fuconate to a 6-carbon acid tentatively identified as 2,4(or 5)-diketo-5(or 4)-monohydroxyhexanoate. The enzyme has a pH optimum of 10. 5 or higher. It is stabilized by dithiothereitol and inhibited by p-hydroxymercuribenzoate and heavy metals (Ag+, Hg2+, Co2+, Cd2+, Pb2+, Zn2+, and Cu2+), suggesting the presence of a functionally essential sulfhydryl group; pre-treatment of enzyme with NAD+ prevents inhibition by p-hydrocymercuribenzoate and heavy metals indicating that this sulfhydryl group may be near the NAD+ binding site. The enzyme has an absolute requirement for NAD+; NADP+ is an ineffective coenzyme. Several lines of evidence indicate that the same enzyme acts on both 2-keto-3-deocy-L-fuconate and 2-keto-3-deoxy-D-arabonate; thus, the pure enzyme acts on both substrates, the two substrates have very similar kinetic parameters (Km values are: 2-keto-3-deocy-L-fuconate, 0.20 mM; 2-keto-3-deoxy-D-arabonate, 0.25 mM; NAD+ for either substrate, 0.22 to 0.25 mM), the two substrates show identical pH and temperature profiles and the two substrates compete for common enzyme active sites. A large number of other sugars and sugar acids, including several 2-keto-3-deoxyaldonates, were ineffective as substrates. The dehydrogenase was also found in calf, beef, lamb, mouse, and rat liver. These studies when considered together with previous studies on the metabolism of L-fucose in pork liver indicate the presence of a soluble enzyme pathway capable of converting L-fucose to 2,4(or 5)-diketo-5(or 4)-monohydroxyhexanoate; this pathway can also convert D-arabinose, and probably L-galactose, to the analogous derivatives (diketomonohydroxypentanoate and diketodihydroxyhexanoate, respectively.  相似文献   

15.
The oxidation of formaldehyde by rat liver mitochondria in the presence of 50 mM phosphate was enhanced 2-fold by exogenous NAD+. Absolute requirement of NAD+ for formaldehyde oxidation was demonstrated by depleting the mitochondria of their NAD+ content (4.6 nmol/mg of protein), followed by reincorporation of the NAD+ into the depleted mitochondria. Aldehyde (formaldehyde) dehydrogenase activity was completely abolished in the depleted mitochondria, but the enzyme activity was restored to control levels following reincorporation of the pyridine nucleotide. Phosphate stimulation of formaldehyde oxidation could not be explained fully by the phosphate-induced swelling which enhances membrane permeability to NAD+, since stimulation of the enzyme activity by increased phosphate concentrations was still observed in the absence of exogenous NAD+. The Km for formaldehyde oxidation by the mitochondria was found to be 0.38 nM, a value similar to that obtained with varying concentrations of NAD+; both Vmax values were very similar, giving a value of 70 to 80 nmol/min/mg of protein. The pH optimum for the mitochondrial enzyme was 8.0. Inhibition of the enzyme activity by anaerobiosis was apparently due to the inability of the respiratory chain to oxidize the generated NADH. The inhibition of mitochondrial formaldehyde oxidation by succinate was found to be due to a lowering of the NAD+ level in the mitochondria. Succinate also inhibited acetaldehyde oxidation by the mitochondria. Malonate, a competitive inhibitor of succinic dehydrogenase, blocked the inhibitory effect of succinate. The respiratory chain inhibitors, rotenone, and antimycin A plus succinate, strongly inhibited formaldehyde oxidation by apparently the same mechanism, although the crude enzyme preparation (freed from the membrane) was slightly sensitive to rotenone. The mitochondria were subfractionated, and 85% of the enzyme activity was found in the inner membrane fraction (mitoplast). Furthermore, separation into inner membrane and matrix components indicated a distribution of aldehyde dehydrogenase activity similar to malic dehydrogenase.  相似文献   

16.
Formaldehyde dehydrogenase was isolated and purified in an overall yield of 12% from cell-free extract of Pseudomonas putida C-83 by chromatographies on columns of DEAE-cellulose, DEAE-Sephadex A-50, and hydroxyapatite. The purified enzyme was homogeneous as judged by disc gel electrophoresis and was most active at pH 7.8 using formaldehyde as a substrate. The enzyme was also active toward acetaldehyde, propionaldehyde, glyoxal, and pyruvaldehyde, though the reaction rates were low. The enzyme was NAD+-linked but did not require the external addition of glutathione, in contrast with the usual formaldehyde dehydrogenase from liver mitochondria, baker's yeast, and some bacteria. The enzyme was markedly inhibited by Ni2+, Pd2+, Hg2+, p-chloromercuribenzoate, and phenylmethanesulfonyl fluoride. The molecular weight of the enzyme was estimated to be 150,000 by the gel filtration method, and analysis by SDS-polyacrylamide gel electrophoresis indicated that the enzyme was composed of two subunit monomers. Kinetic analysis gave Km values of 67 microM for formaldehyde and 56 microM for NAD+, and suggested that the reaction proceeds by a "Ping-pong" mechanism. The enzyme catalyzed the oxidation of formaldehyde accompanied by the stoichiometric reduction of NAD+, but no reverse reaction was observed.  相似文献   

17.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH) catalyzes the oxidation of the 15(S) hydroxyl group of prostaglandins to a 15-keto group resulting in a significant reduction of the biological activities of prostaglandins. Although the key residues involved in NAD+ binding and in catalytic activity have been partially identified, the sites of interaction of the enzyme with the prostaglandin substrates are yet to be determined. Homology analysis of the primary structures of 15-PGDH from human, mouse and rat indicates that the sequences are almost homologous except for two regions near the C-terminus. The involvement of the C-terminal region in catalytic activity was examined by studies on C-terminally truncated enzymes and on human/rat chimeric enzymes. When three to four amino acids were removed successively from the C-terminal end of human 15-PGDH, the truncated enzymes exhibited decreasing Vmax/Km ratios and increasing Km values for PGE2 as the chain was shortened. Similarly, when the C-terminal 14 amino acids of human 15-PGDH were replaced by the C-terminal 14 amino acids of rat 15-PGDH or vice versa, the Vmax/Km ratios and the Km values for prostaglandin E2 of the chimeric enzymes were in between those of the two wild-type enzymes. This indicates that the catalytic effectiveness of human 15-PGDH decreases as the C-terminal region is gradually removed or replaced by rat sequences. The C-terminal region appears to be more important for the interaction of the enzyme with the prostaglandin substrates than with the coenzyme.  相似文献   

18.
The crystalline formate dehydrogenase from Candida methanolica, which showed the highest specific activity (7.52 U/mg) so far reported, was characterized in detail. The enzyme is a dimer composed of identical subunits, each containing one SH group related to the catalytic activity. The molecular mass of the enzyme is about 82-86 kDa. The Km values were found to be 3.0 mM for formate and 0.11 mM for NAD+. Even if the enzyme was incubated at pH 6.5-9.5 or at 55 degrees C, the activity remained at 100%. Hg2+, Ni2+, NaCN, NaN3 and p-chloromercuribenzoate strongly inhibited the enzyme activity, while the enzyme showed relatively high resistance to various chelating agents. The amino acid composition and some other physicochemical properties of the enzyme were studied. Immunological studies revealed that formate dehydrogenases of methanol-utilizing yeasts immunologically more or less resemble each other, but differ from those of methanol-utilizing bacteria. Furthermore, yeast formate dehydrogenases can be immunologically classified into three types: (a) the Candida type, (b) the Torulopis/Hansenula/Pichia type and (c) the formaldehyde-resistant yeast type. For simple and large-scale preparation of the enzyme for practical use, treatment of cells of C. methanolica with the commercial cationic detergent, 'Benzalkonium' cation, is useful: the total and specific activities of the enzyme are 1.17-fold and 3.10-fold higher than those of the crude cell-free extract, respectively.  相似文献   

19.
Freshly prepared samples of yeast alcohol dehydrogenase (EC 1.1.1.1) were inhibited by 1,10-phenanthroline at pH 7.0 and 0 degrees C in a two-stage process. The first step appeared to be slowly established, but was rendered reversible by removal of reagent or by addition of excess Zn2+ ions. The second step was irreversible and was associated with the dissociation of the tetrameric enzyme. The presence of saturating concentrations of NAD+ or NADH promoted and enhanced inhibition by the slowly established reversible process, but prevented dissociation of the enzyme. For the incubation mixtures containing NAD+, removal of the 1,10-phenanthroline resulted in virtually complete recovery of activity, whereas, for the incubation mixtures containing NADH, removal of the reagent gave only partial re-activation. The presence of NAD+ and pyrazole, or NADH and acetamide, in incubation mixtures with the enzyme gave rise to ternary complexes that gave protection against both forms of inactivation by 1,10-phenanthroline. The results support the view that at least some of the Zn2+ ions associated with yeast alcohol dehydrogenase have a catalytic, as opposed to a purely structural, role.  相似文献   

20.
L D Byers  D E Koshland 《Biochemistry》1975,14(16):3661-3669
The specificity of induced conformational changes and of the probes used to detect them has been investigated in yeast glyceraldehyde-3-phosphate dehydrogenase. Cyanylation of the active-site SH groups in two of the four identical subunits of glyceraldehyde-3-phosphate dehydrogenase has no effect on reactivity of the unmodified SH groups toward the cyanylating reagent (2-nitro-5-thiocyanogenzoic acid, NTCB) but results in total loss of catalytic activity. Cyanylation of the dicarboxamidomethylated enzyme was four orders of magnitude slower than with the unmodified enzyme in contrast to cyanylation of the dicyanylated enzyme. Cyanylation by NTCB as well as alkylation by iodoacetate and acylation with beta-(2-furyl)acryloyl phosphate are enhanced in the presence of NAD+ while alkylation by iodoacetamide is inhibited by NAD+. In the absence of NAD+, hydrolysis of the acylated enzyme is faster than phosphorolysis while the reverse is true in the presence of NAD+. NAD+ accelerates hydrolysis of the 3-phosphoglyceroylated enzyme about 60-fold but decreases the rate of hydrolysis of the furylacryloylated enzyme by a factor of 17. Other examples of the specificity of the induced conformational changes and the probes are described. The conformational changes induced by NAD+ make the protein specifically reactive toward its physiological substrates and less reactive toward extraneous competing compounds.  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