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.
Concentrated urine formation in the kidney is accompanied by conditions that favor the accumulation of reactive oxygen species (ROS). Under hyperosmotic conditions, medulla cells accumulate glycine betaine, which is an osmolyte synthesized by betaine aldehyde dehydrogenase (BADH, EC 1.2.1.8). All BADHs identified to date have a highly reactive cysteine residue at the active site, and this cysteine is susceptible to oxidation by hydrogen peroxide. Porcine kidney BADH incubated with H(2)O(2) (0-500 μM) lost 25% of its activity. However, pkBADH inactivation by hydrogen peroxide was limited, even after 120 min of incubation. The presence of coenzyme NAD(+) (10-50 μM) increased the extent of inactivation (60%) at 120 min of reaction, but the ligands betaine aldehyde (50 and 500 μM) and glycine betaine (100 mM) did not change the rate or extent of inactivation as compared to the reaction without ligand. 2-Mercaptoethanol and dithiothreitol, but not reduced glutathione, were able to restore enzyme activity. Mass spectrometry analysis of hydrogen peroxide inactivated BADH revealed oxidation of M278, M243, M241 and H335 in the absence and oxidation of M94, M327 and M278 in the presence of NAD(+). Molecular modeling of BADH revealed that the oxidized methionine and histidine residues are near the NAD(+) binding site. In the presence of the coenzyme, these oxidized residues are proximal to the betaine aldehyde binding site. None of the oxidized amino acid residues participates directly in catalysis. We suggest that pkBADH inactivation by hydrogen peroxide occurs via disulfide bond formation between vicinal catalytic cysteines (C288 and C289).  相似文献   

2.
L Liu  W P Huskey 《Biochemistry》1992,31(30):6898-6903
Primary hydrogen isotope effects and steady-state kinetics have been used to study the mechanism of glyceraldehyde-3-phosphate (GAP) dehydrogenase at pH 8.6. The isotope effect determined by using GAP-1d was unity and independent of arsenate (used as the acyl acceptor) and NAD+ concentrations when the aldehyde substrate was at saturating concentrations. At low GAP concentrations (apparent V/K conditions), the primary hydrogen isotope effect (H/D) was in the range of 1.40-1.52 and independent of arsenate and NAD+ concentrations. Apparent V/K for NAD+ was independent of GAP concentration, and apparent V/K for GAP was independent of NAD+ concentration. The dependence of apparent V/K for GAP on arsenate concentration was more complex but extrapolated to nonzero V/K at the zero-arsenate intercept. These observations are consistent with the general features of the Segal and Boyer (1953) mechanism for the reaction.  相似文献   

3.
In rat adrenal gland and gastric mucosa putrescine is efficiently oxidized to GABA via gamma-aminobutyraldehyde (ABAL) by action of diamine oxidase and aldehyde dehydrogenase. Having turned our attention on the rat intestinal mucosa, where putrescine uptake and diamine oxidase are active, we have purified and characterized an aldehyde dehydrogenase optimally active on gamma-aminobutyraldehyde. A dimer with a subunit molecular weight of 52,000, the native enzyme binds ABAL and NAD+ with high affinity: at pH 7.4, Km values are equal to 18 and 14 microM, respectively. Affinity for betaine aldehyde is much lower (Km = 285 microM), but the efficiency is equally good, thanks to a high value of V. Unaffected by disulfiram and Mg2+, the enzyme is activated by high NAD+ concentrations (Vnn = 1.6 x Vn) and is competitively inhibited by NADH. According to the best fitting model, the dimeric enzyme only binds one NADH and the mixed complex enzyme-NAD(+)-NADH is inactive. The increase of activity promoted by NAD+ can therefore be ascribed to an allosteric effect, rather than to the activation of a second reaction center. Highly stable at pH 6.8 in the presence of dithiothreitol and high phosphate concentrations, ABALDH is inactivated by ion-exchange resins and by cationic buffers. Our results show that the enzyme can be effectively involved in the metabolism of biogenic amines and, with a K(m) for ABAL lower than 20 microM, in the synthesis of GABA.  相似文献   

4.
Porcine kidney betaine aldehyde dehydrogenase (EC 1.2.1.8) kinetic properties were determined at low substrate concentrations. The double-reciprocal plots of initial velocity versus substrate concentration are linear and intersect at the left of the 1/v axis and showed substrate inhibition with betaine aldehyde. Studies of inhibition by NADH and dead-end analogs showed that NADH is a mixed inhibitor against NAD(+) and betaine aldehyde. AMP is competitive with respect to NAD(+) and mixed with betaine aldehyde. Choline is competitive against betaine aldehyde and uncompetitive with respect to NAD(+). The kinetic behavior is consistent with an Iso-Ordered Bi-Bi Steady-State mechanism.  相似文献   

5.
A new assay procedure for measurement of rat liver mitochondrial choline dehydrogenase was developed. Oxidation of [methyl-14C]choline to [methyl-14C]betaine aldehyde and [methyl-14C]betaine was measured after isolating these compounds using HPLC. We observed that NAD+ was required for conversion of betaine aldehyde to betaine in rat liver mitochondria. In the absence of this cofactor, oxidation of choline led to the accumulation of betaine aldehyde. The apparent Km of the mitochondrial choline dehydrogenase for choline was 0.14-0.27 mM, which is significantly lower than previously reported. A partially purified preparation of choline dehydrogenase catalyzed betaine aldehyde formation only in the presence of exogenous electron acceptors (e.g., phenazine methosulfate). This preparation failed to catalyze the formation of betaine even in the presence of NAD+, indicating that betaine aldehyde dehydrogenase may be a separate enzyme from choline dehydrogenase.  相似文献   

6.
Glycine betaine and its precursors choline and glycine betaine aldehyde have been found to confer a high level of osmotic tolerance when added exogenously to cultures of Escherichia coli at an inhibitory osmotic strength. In this paper, the following findings are described. Choline works as an osmoprotectant only under aerobic conditions, whereas glycine betaine aldehyde and glycine betaine function both aerobically and anaerobically. No endogenous glycine betaine accumulation was detectable in osmotically stressed cells grown in the absence of the osmoprotectant itself or the precursors. A membrane-bound, O2-dependent, and electron transfer-linked dehydrogenase was found which oxidized choline to glycine betaine aldehyde and aldehyde to glycine betaine at nearly the same rate. It displayed Michaelis-Menten kinetics; the apparent Km values for choline and glycine betaine aldehyde were 1.5 and 1.6 mM, respectively. Also, a soluble, NAD-dependent dehydrogenase oxidized glycine betaine aldehyde. It displayed Michaelis-Menten kinetics; the apparent Km values for the aldehyde, NAD, and NADP were 0.13, 0.06, and 0.5 mM, respectively. The choline-glycine betaine pathway was osmotically regulated, i.e., full enzymic activities were found only in cells grown aerobically in choline-containing medium at an elevated osmotic strength. Chloramphenicol inhibited the formation of the pathway in osmotically stressed cells.  相似文献   

7.
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   

8.
The kinetics of the NAD+-dependent oxidation of aldehydes, catalysed by aldehyde dehydrogenase purified from sheep liver mitochondria, were studied in detail. Lag phases were observed in the assays, the length of which were dependent on the enzyme concentration. The measured rates after the lag phase was over were directly proportional to the enzyme concentration. If enzyme was preincubated with NAD+, the lag phase was eliminated. Double-reciprocal plots with aldehyde as the variable substrate were non-linear, showing marked substrate activation. With NAD+ as the variable substrate, double-reciprocal plots were linear, and apparently parallel. Double-reciprocal plots with enzyme modified with disulfiram (tetraethylthiuram disulphide) or iodoacetamide, such that at pH 8.0 the activity was decreased to 50% of the control value, showed no substrate activation, and the plots were linear. At pH 7.0, the kinetic parameters Vmax. and Km NAD+- for the oxidation of acetaldehyde and butyraldehyde by the native enzyme are almost identical. Formaldehyde and propionaldehyde show the same apparent maximum rate. Aldehyde dehydrogenase is able to catalyse the hydrolysis of p-nitrophenyl esters. This esterase activity was stimulated by both NAD+ and NADH, the maximum rate for the NAD+ stimulated esterase reaction being roughly equal to the maximum rate for the oxidation of aldehydes. The mechanistic implications of the above behaviour are discussed.  相似文献   

9.
Acetaldehyde metabolism in human erythrocytes was studied using head-space gas chromatographic determination methods, and it was found that acetaldehyde is metabolized in erythrocytes by NAD dependent cytosolic enzyme having an apparent Km value for acetaldehyde approximately 0.7 mM at pH 7.4, and more than 50% of this activity was reduced by 1 μM disulfiram. So, it is suggested that erythrocytes may have an enzyme system similar to the high Km isozyme of the liver aldehyde dehydrogenase.  相似文献   

10.
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   

11.
The osmoregulatory NAD-dependent betaine aldehyde dehydrogenase (betaine aldehyde:NAD oxidoreductase, EC 1.2.1.8), of Escherichia coli, was purified to apparent homogeneity from an over-producing strain carrying the structural gene for the enzyme (betB) on the plasmid vector pBR322. Purification was achieved by ammonium sulfate fractionation of disrupted cells, followed by affinity chromatography on 5'-AMP Sepharose, gel-filtration and ion-exchange chromatography. The amino acid composition was determined. The dehydrogenase was found to be a tetramer with identical 55 kDa subunits. Both NAD and NADP could be used as cofactor for the dehydrogenase, but NAD was preferred. The dehydrogenase was highly specific for betaine aldehyde. None of the analogs tested functioned as a substrate, but several inhibited the enzyme competitively. The enzyme was not activated by salts at concentrations encountered during osmotic upshock, but it was salt tolerant, retaining 50% of maximal activity at 1.2 M K+. It is inferred that salt tolerance is an essential property for an enzyme participating in the cellular synthesis of an osmoprotectant.  相似文献   

12.
Thermal unfolding parameters were determined for a two-domain tetrameric enzyme, phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and for its isolated NAD(+)-binding domain. At pH 8.0, the transition temperatures (t(max)) for the apoforms of the native Bacillus stearothermophilus GAPDH and the isolated domain were 78.3 degrees C and 61.9 degrees C, with calorimetric enthalpies (DeltaH(cal)) of 4415 and 437 kJ/mol (or 30.7 and 22.1 J/g), respectively. In the presence of nearly saturating NAD(+) concentrations, the t(max) and the DeltaH(cal) increased by 13.6 degrees C and by 2365 kJ/mol, respectively, for the native apoenzyme, and by 2.8 degrees C and 109 kJ/mol for the isolated domain. These results indicate that interdomain interactions are essential for NAD(+) to produce its stabilizing effect on the structure of the native enzyme. The thermal stability of the isolated NAD(+)-binding domain increased considerably upon transition from pH 6.0 to 8.0. By contrast, native GAPDH exhibited greater stability at pH 6.0; similar pH-dependencies of thermal stability were displayed by GAPDHs isolated from rabbit muscle and Escherichia coli. The binding of NAD(+) to rabbit muscle apoenzyme increased t(max) and DeltaH(cal) and diminished the widths of the DSC curves; the effect was found to grow progressively with increasing coenzyme concentrations. Alkylation of the essential Cys149 with iodoacetamide destabilized the apoenzyme and altered the effect of NAD(+). Replacement of Cys149 by Ser or by Ala in the B. stearothermophilus GAPDH produced some stabilization, the effect of added NAD(+) being basically similar to that observed with the wild-type enzyme. These data indicate that neither the ion pairing between Cys149 and His176 nor the charge transfer interaction between Cys149 and NAD(+) make any significant contribution to the stabilization of the enzyme's native tertiary structure and the accomplishment of NAD(+)-induced conformational changes. The H176N mutant exhibited dramatically lower heat stability, as reflected in the values of both DeltaH(cal) and t(max). Interestingly, NAD(+) binding resulted in much wider heat capacity curves, suggesting diminished cooperativity of the unfolding transition.  相似文献   

13.
Spinach (Spinacia oleracea L.) leaves contain a nuclear-encoded chloroplastic betaine aldehyde dehydrogenase (EC 1.2.1.8) which is induced several-fold by salinization. Betaine aldehyde dehydrogenase was purified 2400-fold to homogeneity with an overall yield of 14%. The procedure included fractional precipitation with ammonium sulfate, followed by ion-exchange, hydrophobic interaction, and hydroxyapatite chromatography in open columns, and ion-exchange and hydrophobic interaction chromatography in a fast-protein liquid chromatography system. The betaine aldehyde dehydrogenase had a pI of 5.65, and a broad pH optimum between 7.5 and 9.5. The Km values for NAD+ and NADP+ were 20 and 320 microM, respectively; the Vmax of the reaction with NADP+ was 75% of that with NAD+. The native enzyme is a dimer with subunits of Mr 63,000. Highly specific antiserum was raised against the native enzyme, and was used in conjunction with cell-free translation of leaf poly(A)+ RNA to show (a) that betaine aldehyde dehydrogenase is synthesized as a precursor of Mr 1200 higher than the mature polypeptide, and (b) that both chronic salt stress and salt shock provoke a several-fold increase in the level of translatable message for the enzyme.  相似文献   

14.
Choline dehydrogenase (CD), the first enzyme of the glycine betaine synthetic pathway, was measured in a mitochondrial lysate from gill tissue from Atlantic and Chesapeake Bay oysters acclimated to both 350 and 750 mosm. CD from both populations functions at its maximum rate at 30 degrees C and pH 8.75. Although CD from both populations has a similar affinity for its substrate, choline (K(m) = 15.7 mM), CD V(max) from Atlantic oysters is twice that from Bay oysters. In addition, the CD K(m )doubles and the V(max) increases four-fold in both oyster populations acclimated to 750 mosm. CD activity is competitively inhibited by both betaine aldehyde and glycine betaine. The differences in CD kinetics between the two oyster populations help to account for the lower glycine betaine synthesis rates and concentrations in Chesapeake Bay oysters. CD cannot function rapidly enough to saturate the enzyme, betaine aldehyde dehydrogenase (BADH), immediately downstream, and, therefore, CD kinetics limit the rate of glycine betaine synthesis in oysters. J. Exp. Zool. 286:250-261, 2000.  相似文献   

15.
The osmoregulatory NAD-dependent betaine aldehyde dehydrogenase (betaine aldehyde: NAD oxidoreductase, EC 1.2.1.8), of Escherichia coli, was purified to apparent homogeneity from an over-producing strain carrying the structural gene for the enzyme (betB) on the plasmid vector pBR322. Purification was achieved by ammonium sulfate fractionation of disrupted cells, followed by affinity chromatography on 5′-AMP Sepharose, gel-filtration and ion-exchange chromatography. The amino acid composition was determined. The dehydrogenase was found to be a tetramer with identical 55 kDa subunits. Both NAD and NADP could be used as cofactor for the dehydrogenase, but NAD was preferred. The dehydrogenase was highly specific for betaine aldehyde. None of the analogs tested functioned as a substrate, but several inhibited the enzyme competitively. The enzyme was not activated by salts at concentrations encountered during osmotic upshock, but it was salt tolerant, retaining 50% of maximal activity at 1.2 M K+. It is inferred that salt tolerance is an essential property for an enzyme participating in the cellular synthesis of an osmoprotectant.  相似文献   

16.
Three alcohol dehydrogenases have been identified in Acinetobacter calcoaceticus sp. strain HO1-N: an NAD(+)-dependent enzyme and two NADP(+)-dependent enzymes. One of the NADP(+)-dependent alcohol dehydrogenases was partially purified and was specific for long-chain substrates. With tetradecanol as substrate an apparent Km value of 5.2 microM was calculated. This enzyme has a pI of 4.5 and a molecular mass of 144 kDa. All three alcohol dehydrogenases were constitutively expressed. Three aldehyde dehydrogenases were also identified: an NAD(+)-dependent enzyme, an NADP(+)-dependent enzyme and one which was nucleotide independent. The NAD(+)-dependent enzyme represented only 2% of the total activity and was not studied further. The NADP(+)-dependent enzyme was strongly induced by growth of cells on alkanes and was associated with hydrocarbon vesicles. With tetradecanal as substrate an apparent Km value of 0.2 microM was calculated. The nucleotide-independent aldehyde dehydrogenase could use either Würster's Blue or phenazine methosulphate (PMS) as an artificial electron acceptor. This enzyme represents approximately 80% of the total long-chain aldehyde oxidizing activity within the cell when the enzymes were induced by growing the cells on hexadecane. It is particulate but can be solubilized using Triton X-100. The enzyme has an apparent Km of 0.36 mM for decanal.  相似文献   

17.
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   

18.
1. Mitochondrial aldehyde dehydrogenase is purified to near homogeneity by hydroxylapatite-, affinity- and hydrophobic interaction-chromatography. 2. The enzyme is an oligomeric protein and its molecular weight, as determined by gel-filtration, is 117,000 +/- 5000. 3. Active only in the presence of exogenous sulfhydryl compounds and NAD(+)-dependent, aldehyde dehydrogenase works optimally with linear-chain aliphatic aldehydes and is practically inactive with benzaldehyde. The pH-optimum is at about pH 8.5. 4. Km-Values for aliphatic aldehydes (C2-C6) range between 0.17 and 0.32 microM. The Km for NAD+ increases from 16 microM with acetaldehyde to 71 microM with capronaldehyde. 5. Millimolar concentrations of Mg2+ promote high increases of both V and Km for NAD+. At the same time, saturation curves with C4-C6 aldehydes can be simulated with a substrate inhibition model. 6. Inhibition by NADH is competitive: with capronaldehyde, the inhibition constant for NADH is 52 microM in the absence of Mg2+ and 14 microM in the presence of 4 mM Mg2+; with acetaldehyde, the inhibition constant is about three times higher (36 and 159 microM, respectively).  相似文献   

19.
An NAD+-linked 17 beta-hydroxysteroid dehydrogenase was purified to homogeneity from a fungus, Cylindrocarpon radicicola ATCC 11011 by ion exchange, gel filtration, and hydrophobic chromatographies. The purified preparation of the dehydrogenase showed an apparent molecular weight of 58,600 by gel filtration and polyacrylamide gel electrophoresis. SDS-gel electrophoresis gave Mr = 26,000 for the identical subunits of the protein. The amino-terminal residue of the enzyme protein was determined to be glycine. The enzyme catalyzed the oxidation of 17 beta-hydroxysteroids to the ketosteroids with the reduction of NAD+, which was a specific hydrogen acceptor, and also catalyzed the reduction of 17-ketosteroids with the consumption of NADH. The optimum pH of the dehydrogenase reaction was 10 and that of the reductase reaction was 7.0. The enzyme had a high specific activity for the oxidation of testosterone (Vmax = 85 mumol/min/mg; Km for the steroid = 9.5 microM; Km for NAD+ = 198 microM at pH 10.0) and for the reduction of androstenedione (Vmax = 1.8 mumol/min/mg; Km for the steroid = 24 microM; Km for NADH = 6.8 microM at pH 7.0). In the purified enzyme preparation, no activity of 3 alpha-hydroxysteroid dehydrogenase, 3 beta-hydroxysteroid dehydrogenase, delta 5-3-ketosteroid-4,5-isomerase, or steroid ring A-delta-dehydrogenase was detected. Among several steroids tested, only 17 beta-hydroxysteroids such as testosterone, estradiol-17 beta, and 11 beta-hydroxytestosterone, were oxidized, indicating that the enzyme has a high specificity for the substrate steroid. The stereospecificity of hydrogen transfer by the enzyme in dehydrogenation was examined with [17 alpha-3H]testosterone.  相似文献   

20.
Fan F  Germann MW  Gadda G 《Biochemistry》2006,45(6):1979-1986
Choline oxidase catalyzes the four-electron oxidation of choline to glycine betaine via two sequential FAD-dependent reactions in which betaine aldehyde is formed as an intermediate. The chemical mechanism for the oxidation of choline catalyzed by choline oxidase was recently elucidated by using kinetic isotope effects [Fan, F., and Gadda, G. (2005) J. Am. Chem. Soc. 127, 2067-2074]. In this study, the oxidation of betaine aldehyde has been investigated by using spectroscopic and kinetic analyses with betaine aldehyde and its isosteric analogue 3,3-dimethylbutyraldehyde. The pH dependence of the kcat/Km and kcat values with betaine aldehyde showed that a catalytic base with a pKa of approximately 6.7 is required for betaine aldehyde oxidation. Complete reduction of the enzyme-bound flavin was observed in a stopped-flow spectrophotometer upon anaerobic mixing with betaine aldehyde or choline at pH 8, with similar k(red) values > or = 48 s(-1). In contrast, only 10-26% of the enzyme-bound flavin was reduced by 3,3-dimethylbutyraldehyde between pH 6 and 10. Furthermore, this compound acted as a competitive inhibitor versus choline. NMR spectroscopic analyses indicated that betaine aldehyde exists predominantly (99%) as a diol form in aqueous solution. In contrast, the thermodynamic equilibrium for 3,3-dimethylbutyraldehyde favors the aldehyde (> or = 65%) over the hydrated form in the pH range from 6 to 10. The keto species of 3,3-dimethylbutyraldehyde is reactive toward enzymic nucleophiles, as suggested by the kinetic data with NAD+-dependent yeast aldehyde dehydrogenase. The data presented suggest that choline oxidase utilizes the hydrated species of the aldehyde as substrate in a mechanism for aldehyde oxidation in which hydride transfer is triggered by an active site base.  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