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 31 毫秒
1.
1. Washed-cell suspensions of Escherichia coli, incubated at the optimum pH of 6.4 and with a saturating substrate concentration of approx. 10mm, convert dl-1-aminopropan-2-ol into aminoacetone at a rate of approx. 4.0mmumoles/mg. dry wt. of cells/min. at 30 degrees . 2. Mg(2+), Mn(2+), Co(2+), Zn(2+), Ca(2+), K(+) and NH(4) (+), as sulphates, and EDTA have no effect on this rate, although Cu(2+) inhibits and Fe(2+) activates to some extent. 3. Conditions of growth markedly affect the rate of aminoacetone production by cell suspensions. 4. Dialysed cell-free extracts of E. coli exhibit 1-aminopropan-2-ol-dehydrogenase activity, the enzyme having optimum activity at pH7.0, a requirement for NAD(+) and K(+), and a K(m) for the amino alcohol substrate of 0.8mm, calculated for a single enantiomorph. 5. Under optimum conditions 1-aminopropan-2-ol dehydrogenase forms aminoacetone at rate of approx. 3.0mmumoles/mg. of protein/min. at 37 degrees . The enzyme is only slightly inhibited by dl-3-hydroxybutyrate and dl-2-hydroxy-2-phenylethyl-amine. 6. l-Threonine-dehydrogenase activity is exhibited by both whole cells and cell-free extracts. Whole cells produce aminoacetone from l-threonine more slowly than they do from dl-1-aminopropan-2-ol, whereas the situation is reversed in cell-free extracts. Both kinetic evidence, and the fact that synthesis of 1-aminopropan-2-ol dehydrogenase, but not of threonine dehydrogenase, is repressed by compounds such as glucose and pyruvate, provide evidence that the amino alcohol is oxidized by a specific enyme. 7. The metabolic role of 1-aminopropan-2-ol dehydrogenase is discussed.  相似文献   

2.
1. Pseudomonas sp. N.C.I.B. 8858 grew well on d- and l-1-aminopropan-2-ol and on aminoacetone. 2. Cell-free extracts possessed high activities of inducibly formed l-1-aminopropan-2-ol-NAD(+) oxidoreductase, amino alcohol-ATP phosphotransferase, dl-1-aminopropan-2-ol O-phosphate phospho-lyase and aldehyde-NAD(+) oxidoreductase, but no 1-aminopropan-2-ol racemase or d-1-aminopropan-2-ol-NAD(+) oxidoreductase. 3. The amino alcohol kinase (activated by ADP) was non-stereospecific towards 1-aminopropan-2-ol and was one-third as active with ethanolamine. The phospho-lyase was active with l- and d-1-aminopropan-2-ol O-phosphate, but ethanolamine O-phosphate was only one-tenth as active as its higher homologues. The purified aldehyde dehydrogenase was active with propionaldehyde, acetaldehyde and also with methylglyoxal. The previously observed 2-oxo aldehyde dehydrogenase activity was considered to be due to the broadly specific aldehyde dehydrogenase. 4. Mutants of Pseudomonas sp. N.C.I.B. 8858 deficient in 1-aminopropan-2-ol kinase, 1-aminopropan-2-ol O-phosphate phospho-lyase, aldehyde dehydrogenase or an enzyme involved in propionate metabolism were incapable of growth on aminoacetone or 1-aminopropan-2-ol as carbon source, although all except the kinase- or phospho-lyasedeficient mutants could use these compounds and ethanolamine as nitrogen sources. The aldehyde dehydrogenase-deficient mutants produced copious amounts of propionaldehyde and acetaldehyde during growth on the corresponding amino alcohols. 5. The path of aminoacetone metabolism in Pseudomonas sp. N.C.I.B. 8858 was concluded to involve l-1-aminopropan-2-ol, the O-phosphate ester of this compound, propionaldehyde and propionate as obligatory intermediates. d-1-Aminopropan-2-ol was metabolized by the same route as the l-isomer, gratuitously inducing formation of the stereospecific l-1-aminopropan-2-ol dehydrogenase. 6. Extracts of the pseudomonad grown with ethanolamine as the nitrogen source were devoid of 1-aminopropan-2-ol dehydrogenase, the kinase and the phospho-lyase, but exhibited cobamide coenzyme-dependent deaminase activity. Mutants deficient in kinase or phospho-lyase (deaminating) grew well on ethanolamine as the nitrogen source. Ethanolamine deaminase was inactive with, but inhibited by, 1-aminopropan-2-ol.  相似文献   

3.
1. Growth and manometric experiments showed that a Pseudomonas sp. P6 (N.C.I.B. 10431), formerly known as Achromobacter sp. P6, was capable of growth on both stereoisomers of 1-aminopropan-2-ol, and supported the hypothesis that assimilation involved metabolism to propionaldehyde, propionate and possibly 2-hydroxyglutarate. A number of alternative intermediary metabolites were ruled out. 2. Accumulation of propionaldehyde from 1-aminopropan-2-ol by intact cells occurred only during active growth, was transitory and was accompanied by morphological changes in the pseudomonad. 3. Enzymic and radioactive tracer evidence showed that 1-aminopropan-2-ol O-phosphate was the intermediate between amino alcohol and aldehyde. The operation of an inducibly formed ATP-amino alcohol phosphotransferase was established by measuring substrate disappearance, ADP formation and amino alcohol O-phosphate formation. This novel kinase had two activity peaks at about pH7 and 9. It acted on both l- and d-isomers of 1-aminopropan-2-ol, and also on l-threonine and ethanolamine, but had only low activity towards choline. The enzyme was partially purified by ion-exchange chromatography. 4. An amino alcohol O-phosphate phospho-lyase (deaminating) produced propionaldehyde from dl- and d-1-aminopropan-2-ol O-phosphate, and also formed acetaldehyde less rapidly from ethanolamine O-phosphate. It had optimum activity at about pH8 in Tris-HCl buffers. The enzyme was partially purified and evidence was obtained that a single enzyme was responsible for both activities. Apparent K(m) values for the substrates were determined. Activity was inhibited by dl-threonine O-phosphate, dl-serine O-phosphate, choline O-phosphate and P(i). Enzyme formation was induced by growth with either amino alcohol substrate. 5. Radioactive tracer experiments with dl-1-amino[3-(14)C]propan-2-ol confirmed the operation of the amino alcohol kinase and demonstrated coupling with the phospho-lyase enzyme in vitro to produce [(14)C]-propionaldehyde. 6. An aldehyde dehydrogenase, found in extracts of the pseudomonad after growth on 1-aminopropan-2-ol, was characterized and concluded to be responsible for propionaldehyde and acetaldehyde oxidation. The enzyme was inactive with methylglyoxal. 7. Propionate and acetate were concluded to be metabolized via propionyl-CoA and acetyl-CoA, and studies were made of a CoA ester synthase found in extracts. 8. Studies of a strain of Pseudomonas putida N.C.I.B. 10558 suggested that 1-aminopropan-2-ols were metabolized via their O-phosphates, propionaldehyde and propionate. Amino alcohol kinase activity was detected and extracts contained a phospho-lyase showing higher activity with the 1-aminopropan-2-ol O-phosphate than with ethanolamine O-phosphate.  相似文献   

4.
1. A species of Arthrobacter (designated Arthrobacter 9759) was isolated from soil by its ability to grow aerobically on l-threonine as sole source of carbon atoms, nitrogen atoms and energy; the organism also grew well on other sources of carbon atoms including glycine, but no growth was obtainable on aminoacetone or dl-1-aminopropan-2-ol. 2. During growth on threonine, (14)C from l-[U-(14)C]threonine was rapidly incorporated into glycine and citrate, and thereafter into serine, alanine, aspartate and glutamate. 3. With extracts of threonine-grown cells supplied with l-[U-(14)C]threonine, evidence was obtained of the NAD and CoA-dependent catabolism of l-threonine to produce acetyl-CoA plus glycine. Short-term incorporation studies in which [2-(14)C]acetate and [2-(14)C]glycine were supplied (a) to cultures growing on threonine, and (b) to extracts of threonine-grown cells, showed that the acetyl-CoA was metabolized via the tricarboxylic acid cycle and glyoxylate cycle whereas the glycine was converted into pyruvate via the folate-dependent ;serine pathway'. 4. The threonine-grown organism contained ;biosynthetic' threonine dehydratase and a potent NAD-linked l-threonine dehydrogenase but possessed no l-threonine aldolase activity. 5. Evidence was obtained that the acetyl-CoA and glycine produced from l-threonine had their immediate origin in the alpha-amino-beta-oxobutyrate formed by the threonine dehydrogenase; the CoA-dependent cleavage of this compound was catalysed by an alpha-amino-beta-oxobutyrate CoA-ligase, which was identified with ;aminoacetone synthase'. A continuous spectrophotometric assay of this enzyme was developed, and it was found to be inducibly synthesized only during growth on threonine and not during growth on acetate plus glycine. 6. By using a reconstituted mixture of separately purified l-threonine dehydrogenase and alpha-amino-beta-oxobutyrate CoA-ligase (i.e. ;aminoacetone synthase'), l-[U-(14)C]threonine was broken down to [(14)C]glycine plus [(14)C]acetyl-CoA (trapped as [(14)C]citrate). 7. There was no evidence of aminoacetone metabolism by Arthrobacter 9759 even though a small amount of this amino ketone appeared in the culture medium during growth on threonine.  相似文献   

5.
Stopped-flow studies of oxidation of butan-1-ol and propan-2-ol by NAD(+) in the presence of Phenol Red and large concentrations of yeast alcohol dehydrogenase give no evidence for the participation of a group of pK(a) approx. 7.6 in alcohol binding. Such a group has been implicated in ethanol binding to horse liver alcohol dehydrogenase [Shore, Gutfreund, Brooks, Santiago & Santiago (1974) Biochemistry13, 4185-4190]. The present result supports previous findings based on steady-state kinetic studies with the yeast enzyme. Stopped-flow studies of the yeast alcohol dehydrogenase-catalysed reduction of acetaldehyde by NADH in the presence of ethanol as product inhibitor indicate that the rate-limiting step is NAD(+) release from the enzyme-NAD(+)-ethanol product complex. This finding permits calculation of K(3), the dissociation constant for ethanol from the enzyme-NAD(+)-ethanol complex, by using the product-inhibition data of Dickenson & Dickinson (1978) (Biochem. J.171, 613-627). The calculations show that K(3) varies very little with pH in the range 5.95-8.9, and this agrees with the findings of the stopped-flow experiments described above. Absorption and fluorescence measurements on mixtures of substrates and coenzymes in the presence of high concentrations of alcohol dehydrogenase have been used to estimate values for the ratio [enzyme-NADH-acetaldehyde]/ [enzyme-NAD(+)-ethanol] at equilibrium. The values obtained were in the range 0.11+/-0.04, and this value together with estimates of K(3) was used to provide estimates of values for rate constants and dissociation constants for steps within the catalytic mechanism.  相似文献   

6.
1. The 120-fold purification of ethanolamine ammonia-lyase from Escherichia coli extracts, to apparent homogeneity, is described. Ethanolamine, dithiothreitol, glycerol and KCl protected the apoenzyme from inactivation. 2. At the optimum pH7.5, K(m) values for ethanolamine and coenzyme B(12) were 44mum and 0.42mum respectively. The K(m) for ethanolamine was markedly affected by pH, transitions occurring at pH7.0 and 8.35. 3. The enzyme was specific for ethanolamine as substrate, none of the 18 analogues tested being active. l-2-Aminopropan-l-ol (K(i) 0.86mum), dl-1-aminopropan-2-ol (K(i) 2.2mum) and dl-1,3-diaminopropan-2-ol (K(i) 88.0mum) inhibited competitively. 4. Enzyme activity was inhibited, irreversibly and non-competitively, by the coenzyme analogues methylcobalamin (K(i) 1.4nm), hydroxocobalamin (K(i) 2.1nm) and cyanocobalamin (K(i) 4.8nm). 5. Iodoacetamide inhibited in the absence of ethanolamine, but only slightly in its presence. p-Hydroxymercuribenzoate inhibited markedly even in the presence of ethanolamine. Dithiothreitol and 2-mercaptoethanol (less effectively) restored activity to the enzyme dialysed against buffer containing ethanolamine. 6. Although K(+) ions stabilized the enzyme during dialysis or storage, they were not necessary for activity. 7. Gel filtration showed the enzyme to be of high molecular weight, ultracentrifugal studies giving s(20,w) of 16.4 and an estimated mol.wt. 560400. The isoelectric point for the apoenzyme was approx. pH5.0. inhibited enzyme activity at concentrations above 1m (95% inhibition at 3m) and sodium dodecyl sulphate/polyacrylamide-gel electrophoresis indicated protein subunits of mol.wt. 61400. 8. Immunological studies showed that the E.coli enzyme was closely related to those of other enterobacteria, but only distantly to that of Clostridium sp. A double precipitin band suggested that the apoenzyme may be made up of two protein components.  相似文献   

7.
1. Rat tissue homogenates convert dl-1-aminopropan-2-ol into aminoacetone. Liver homogenates have relatively high aminopropanol-dehydrogenase activity compared with kidney, heart, spleen and muscle preparations. 2. Maximum activity of liver homogenates is exhibited at pH9·8. The Km for aminopropanol is approx. 15mm, calculated for a single enantiomorph, and the maximum activity is approx. 9mμmoles of aminoacetone formed/mg. wet wt. of liver/hr.at 37°. Aminoacetone is also formed from l-threonine, but less rapidly. An unidentified amino ketone is formed from dl-4-amino-3-hydroxybutyrate, the Km for which is approx. 200mm at pH9·8. 3. Aminopropanol-dehydrogenase activity in homogenates is inhibited non-competitively by dl-3-hydroxybutyrate, the Ki being approx. 200mm. EDTA and other chelating agents are weakly inhibitory, and whereas potassium chloride activates slightly at low concentrations, inhibition occurs at 50–100mm. 4. It is concluded that aminopropanol-dehydrogenase is located in mitochondria, and in contrast with l-threonine dehydrogenase can be readily solubilized from mitochondrial preparations by ultrasonic treatment. 5. Soluble extracts of disintegrated mitochondria exhibit maximum aminopropanol-dehydrogenase activity at pH9·1 At this pH, Km values for the amino alcohol and NAD+ are approx. 200 and 1·3mm respectively. Under optimum conditions the maximum velocity is approx. 70mμmoles of aminoacetone formed/mg. of protein/hr. at 37°. Chelating agents and thiol reagents appear to have little effect on enzyme activity, but potassium chloride inhibits at all concentrations tested up to 80mm. dl-3-Hydroxybutyrate is only slightly inhibitory. 6. Dehydrogenase activities for l-threonine and dl-4-amino-3-hydroxybutyrate appear to be distinct from that for aminopropanol. 7. Intraperitoneal injection of aminopropanol into rats leads to excretion of aminoacetone in the urine. Aminoacetone excretion proportional to the amount of the amino alcohol administered, is complete within 24hr., but represents less than 0·1% of the dose given. 8. The possible metabolic role of amino alcohol dehydrogenases is discussed.  相似文献   

8.
1. A new induced NAD-dependent hydroaromatic dehydrogenase was isolated from a cell-free extract of Lactobacillus plantarum 13a and purified 175-fold. 2. The enzyme catalyses the oxidation of (-)-quinate, (-)-shikimate, (-)-dihydroshikimate and (-)t-3,t-4-dihydroxycyclohexane-c-1-carboxylate with NAD(+), and the reverse action with NADH. 3. The K(m) values at the optimal pH10.0 for these substrates are 0.85, 0.75, 0.52 and 0.74mm respectively, and the corresponding values for NAD(+) are 0.45, 0.26, 0.34 and 0.36mm respectively. 4. The stereochemical requirements of the enzyme and the role it may play in the reduction of (-)-quinate to (-)t-3,t-4-dihydroxycyclohexane-c-1-carboxylate are discussed and a pathway is suggested.  相似文献   

9.
10.
1. Kinetic experiments suggested the possible existence of at least two different NAD(+)-dependent aldehyde dehydrogenases in rat liver. Distribution studies showed that one enzyme, designated enzyme I, was exclusively localized in the mitochondria and that another enzyme, designated enzyme II, was localized in both the mitochondria and the microsomal fraction. 2. A NADP(+)-dependent enzyme was also found in the mitochondria and the microsomal fraction and it is suggested that this enzyme is identical with enzyme II. 3. The K(m) for acetaldehyde was apparently less than 10mum for enzyme I and 0.9-1.7mm for enzyme II. The K(m) for NAD(+) was similar for both enzymes (20-30mum). The K(m) for NADP(+) was 2-3mm and for acetaldehyde 0.5-0.7mm for the NADP(+)-dependent activity. 4. The NAD(+)-dependent enzymes show pH optima between 9 and 10. The highest activity was found in pyrophosphate buffer for both enzymes. In phosphate buffer there was a striking difference in activity between the two enzymes. Compared with the activity in pyrophosphate buffer, the activity of enzyme II was uninfluenced, whereas the activity of enzyme I was very low. 5. The results are compared with those of earlier investigations on the distribution of aldehyde dehydrogenase and with the results from purified enzymes from different sources.  相似文献   

11.
Abstract A fragment of Methylobacter marinus A45 DNA has been cloned and sequenced, and an open reading frame has been identified that could code for a 46-kDa polypeptide. Comparison of the deduced amino acid sequence of the polypeptide against the protein data bank has revealed strong similarity with a number of alcohol dehydrogenases, with highest similarity towards class III alcohol dehydrogenases, which recently have been shown to be identical to glutathione-dependent formaldehyde dehydrogenases. We were unable to measure appreciable levels of NAD(P)-dependent formaldehyde dehydrogenases or alcohol dehydrogenase activities using aldehydes or primary or secondary alcohols in cell-free extracts from batch cultures of M. marinus A45. However, formaldehyde dehydrogenases activity was detected on zymograms. Our data suggest that, although NAD(P)-linked formaldehyde dehydrogenase or alcohol dehydrogenase activities are undetectable in cell-free extracts of most methylotrophs employing the ribulose monophosphate pathway for formaldehyde assimilation and dissimilation, the gene encoding formaldehyde dehydrogenase is present in M. marinus A45 and may be present in more of these organisms as well.  相似文献   

12.
In the oxidation of methylglyoxal by 2-oxoaldehyde dehydrogenase, the apparent Km value for NADP+ was about 2.5 times lower than the corresponding Km for NAD+; the apparent Km values for methylglyoxal and for the amine activator L-2-aminopropan-1-ol, with NADP+ as cofactor, were also different from those obtained with NAD+. In the presence of NADP+, the enzyme was not activated by P1, in contrast with the activation of the enzyme when NAD+ was used. The significance of the results is discussed.  相似文献   

13.
l-Threonine level in blood plasma is a biomarker of some diseases and nitrogen imbalance in the body. The determination of l-threonine is interesting and is required for diagnosis and management of inherited metabolic disorder. This is the first report of the specific enzymatic determination of l-threonine by a newly discovered l-threonine 3-dehydrogenase (ThrDH, EC 1.1.1.103) from Cupriavidus necator NBRC 102504. ThrDH, a key enzyme in l-threonine catabolism in microorganisms and animals, catalyzes the NAD(+)-dependent oxidation of l-threonine to 2-amino-3-oxobutyrate. ThrDH from C. necator was purified to homogeneity and fully characterized. l-Threonine and dl-2-amino-3-hydroxyvalerate are the only substrates for ThrDH among other l-amino acids, alcohols, and amino alcohols. The primary amino acid structure of ThrDH belongs to the extended short-chain alcohol dehydrogenase superfamily and is related to GDP-mannose-3',5'-epimerase (GME) from Arabidopsis thaliana. Both enzymes have a glycine-rich NAD(+)-binding domain at the N terminal and conserved catalytic triad of YxxxK residues, but substrate-binding residues of GME were not found in the ThrDH sequence. ThrDH significantly differs from known bacterial and archaea ThrDHs that belong to zinc-binding medium chain alcohol dehydrogenase because of low sequence similarity and the lack of a zinc-binding domain in the sequence. A specific, quantitative, and sensitive enzymatic endpoint method for l-threonine determination was developed by using a ThrDH microplate assay. The assay was successfully applied for determination of l-threonine in human serum and plasma. Our specific determination is simple, convenient, inexpensive, accurate, and suitable for mass screening determination of l-threonine in a number of samples.  相似文献   

14.
Multiple forms of lactate dehydrogenase in Staphylococcus aureus   总被引:3,自引:0,他引:3       下载免费PDF全文
Activities for nicotinamide adenine dinucleotide (NAD)-dependent and NAD-independent forms of lactate dehydrogenase (LDH) were measured in cell-free extracts of Staphylococcus aureus strain PS 6 for the d and l isomers of lactate. Data obtained for the NAD-dependent lactate dehydrogenases indicate that oxidation of both isomers of lactate is due to both an l-lactate-specific LDH and a lactate racemase. After acrylamide gel electrophoresis, two bands exhibiting LDH activity were detected in crude or in partially purified cell-free extracts. The fast band exhibited LDH activity that was not NAD-dependent for both isomers of lactate, whereas, the slow band had very high NAD-dependent LDH activity for the l isomer but just detectable activity or the d isomer. Both bands appeared when d-lactate was used as the substrate, but only the slow band was formed when l-lactate was the substrate. NAD-dependent LDH, in apparent association with a nonspecific tetrazolium-reducing protein, is responsible for the production of the slow band.  相似文献   

15.
A mixture of threonine dehydrogenase and aminoacetone synthetase will catalyze the conversion of L-threonine to glycine. The overall reaction likely involves the conversion of L-threonine, NAD+, and CoA to glycine, NADH, and acetyl-CoA. Physical separation of L-threonine dehydrogenase from aminoacetone synthetase results in the formation of aminoacetone and CO2 from their substrates. A physical interaction between threonine dehydrogenase and aminoacetone synthetase has been demonstrated by gel permeation chromatography and fluorescence polarization. Polarization of fluorescence measurements of threonine dehydrogenase and aminoacetone synthetase labeled with fluorescein isothiocyanate indicated the formation of a soluble active complex, with an apparent dissociation constant (Kd) of 5-10 nM and an apparent stoichiometry of 2 aminoacetone synthetase dimers/1 threonine dehydrogenase tetramer. Chemical experiments have identified aminoacetone as the enzymatic product of L-threonine dehydrogenase acting on L-threonine. These experiments involved trapping pyrrole derivatives, [3H]NaBH4 reduction, and coupling with plasma amine oxidase. Kinetic experiments also showed NADH, CO2, and aminoacetone to inhibit threonine dehydrogenase in a manner consistent with an ordered Bi-Ter kinetic mechanism. NAD+ is the lead substrate followed by threonine, and the products are released in the order: CO2, aminoacetone, and NADH.  相似文献   

16.
There is doubt about the l-threonine 3-dehydrogenase (EC 1.1.1.103) and threonine aldolase (EC 2.1.2.1) catabolic pathways of l-threonine in mammals which are believed to produce aminoacetone and glycine plus acetaldehyde, respectively. l-Threonine 3-dehydrogenase in disrupted guinea-pig liver mitochondria was investigated in a reaction mixture containing l-threonine without and with CoA and oxaloacetate; l-[U-14C]threonine was included in four similar experiments for autoradiograms. Threonine aldolase was examined in similar mitochondria from liver and kidney. CoA reduced the aminoacetone formed from l-threonine to 10–14% and CoA plus oxaloacetate produced citrate (from CoASAc) in approximately equal amounts to the decrease in aminoacetone. Autoradiograms confirmed the decrease in aminoacetone with the simultaneous appearance of citrate and glycine. No evidence was obtained that threonine aldolase catabolised l-threonine at the concentration used to assay the dehydrogenase. It is concluded that 2-amino-3-oxobutyrate (precursor of aminoacetone), which is produced from l-threonine by l-threonine 3-dehydrogenase, undergoes CoA-dependent cleavage to glycine and CoASAc by 2-amino-3-oxobutyrate-CoA ligase. The results suggest that the coupling of these enzymes provides a new pathway for the catabolism of threonine in mammals.  相似文献   

17.
Threonine is a precursor of glycine in the rat, but the metabolic pathway involved is unclear. To elucidate this pathway, the biosynthesis of glycine, and of aminoacetone, from l-threonine were studied in rat liver mitochondrial preparations of differing integrities. In the absence of added cofactors, intact mitochondria formed glycine and aminoacetone in approximately equal amounts from 20 mM l-threonine, but exogenous NAD+ decreased and CoA increased the ratio of glycine to aminoacetone formed. In intact and freeze-thawed mitochondria, the ratio of glycine to aminoacetone formed was markedly sensitive to the concentration of l-threonine, glycine being the major product at low l-threonine concentrations. Disruption of mitochondrial integrity by sonication (1 min) decreased the ratio of glycine to aminoacetone formed, and in 20 000 × g supernatant fractions from sonicated (3 min) mitochondria, aminoacetone was the major product. The main non-nitogenous tow-carbon compound detected when intact mitochondria catabolized l-threonine to glycine was acetate, which was probably derived from deacylation of acetyl-CoA. These results suggest that glycine formation from l-threonine in rat liver mitochondria occured primarily by the coupled activities of threonine dehydrogenase and 2-amino-3-oxobutyrate CoA-ligase, the extent of coupling between the enzymes being dependent upon a close physical relationship and upon the flux through the dehydrogenase reaction. In vivo glycine synthesis would predominate, and aminoacetone would be a minor product.  相似文献   

18.
Phosphite dehydrogenase (PTDH) catalyzes the NAD-dependent oxidation of phosphite to phosphate, a reaction that is 15 kcal/mol exergonic. The enzyme belongs to the family of D-hydroxy acid dehydrogenases. Five other family members that were analyzed do not catalyze the oxidation of phosphite, ruling out the possibility that this is a ubiquitous activity of these proteins. PTDH does not accept any alternative substrates such as thiophosphite, hydrated aldehydes, and methylphosphinate, and potential small nucleophiles such as hydroxylamine, fluoride, methanol, and trifluoromethanol do not compete with water in the displacement of the hydride from phosphite. The pH dependence of k(cat)/K(m,phosphite) is bell-shaped with a pK(a) of 6.8 for the acidic limb and a pK(a) of 7.8 for the basic limb. The pK(a) of 6.8 is assigned to the second deprotonation of phosphite. However, whether the dianionic form of phosphite is the true substrate is not clear since a reverse protonation mechanism is also consistent with the available data. Unlike k(cat)/K(m,phosphite), k(cat) and k(cat)/K(m,NAD) are pH-independent. Sulfite is a strong inhibitor of PTDH that is competitive with respect to phosphite and uncompetitive with respect to NAD(+). Incubation of the enzyme with NAD(+) and low concentrations of sulfite results in a covalent adduct between NAD(+) and sulfite in the active site of the enzyme that binds very tightly. Fluorescent titration studies provided the apparent dissociation constants for NAD(+), NADH, sulfite, and the sulfite-NAD(+) adduct. Substrate isotope effect studies with deuterium-labeled phosphite resulted in small normal isotope effects (1.4-2.1) on both k(cat) and k(cat)/K(m,phosphite) at pH 7.25 and 8.0. Solvent isotope effects (SIEs) on k(cat) are similar in size; however, the SIE of k(cat)/K(m,phosphite) at pH 7.25 is significantly larger (4.4), whereas at pH 8.0, it is the inverse (0.6). The pH-rate profile of k(cat)/K(m,phosphite), which predicts that the observed SIEs will have a significant thermodynamic origin, can account for these effects.  相似文献   

19.
There is doubt about the l-threonine 3-dehydrogenase (EC 1.1.1.103) and threonine aldolase (EC 2.1.2.1) catabolic pathways of l-threonine in mammals which are believed to produce aminoacetone and glycine plus acetaldehyde, respectively. l-Threonine 3-dehydrogenase in disrupted guinea-pig liver mitochondria was investigated in a reaction mixture containing l-threonine without and with CoA and oxaloacetate; l-[U-14C]threonine was included in four similar experiments for autoradiograms. Threonine aldolase was examined in similar mitochondria from liver and kidney. CoA reduced the aminoacetone formed from l-threonine to 10–14% and CoA plus oxaloacetate produced citrate (from CoASAc) in approximately equal amounts to the decrease in aminoacetone. Autoradiograms confirmed the decrease in aminoacetone with the simultaneous appearance of citrate and glycine. No evidence was obtained that threonine aldolase catabolised l-threonine at the concentration used to assay the dehydrogenase. It is concluded that 2-amino-3-oxobutyrate (precursor of aminoacetone), which is produced from l-threonine by l-threonine 3-dehydrogenase, undergoes CoA-dependent cleavage to glycine and CoASAc by 2-amino-3-oxobutyrate-CoA ligase. The results suggest that the coupling of these enzymes provides a new pathway for the catabolism of threonine in mammals.  相似文献   

20.
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   

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