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.
Glutaconyl-coenzyme A (CoA) is the presumed enzyme-bound intermediate in the oxidative decarboxylation of glutaryl-CoA that is catalyzed by glutaryl-CoA dehydrogenase. We demonstrated glutaconyl-CoA bound to glutaryl-CoA dehydrogenase after anaerobic reduction of the dehydrogenase with glutaryl-CoA. Glutaryl-CoA dehydrogenase also has intrinsic enoyl-CoA hydratase activity, a property of other members of the acyl-CoA dehydrogenase family. The enzyme rapidly hydrates glutaconyl-CoA at pH 7.6 with a k(cat) of 2.7 s(-1). The k(cat) in the overall oxidation-decarboxylation reaction at pH 7.6 is about 9 s(-1). The binding of glutaconyl-CoA was quantitatively assessed from the K(m) in the hydratase reaction, 3 microM, and the K(i), 1.0 microM, as a competitive inhibitor of the dehydrogenase. These values compare with K(m) and K(i) of 4.0 and 12.9 microM, respectively, for crotonyl-CoA. Glu370 is the general base catalyst in the dehydrogenase that abstracts an alpha-proton of the substrate to initiate the catalytic pathway. The mutant dehydrogenase, Glu370Gln, is inactive in the dehydrogenation and the hydratase reactions. However, this mutant dehydrogenase decarboxylates glutaconyl-CoA to crotonyl-CoA without oxidation-reduction reactions of the dehydrogenase flavin. Addition of glutaconyl-CoA to this mutant dehydrogenase results in a rapid, transient increase in long-wavelength absorbance (lambda(max) approximately 725 nm), and crotonyl-CoA is found as the sole product. We propose that this 725 nm-absorbing species is the delocalized crotonyl-CoA anion that follows decarboxylation and that the decay is the result of slow protonation of the anion in the absence of the general acid catalyst, Glu370(H(+)). In the absence of detectable oxidation-reduction, the data indicate that oxidation-reduction of the dehydrogenase flavin is not essential for decarboxylation of glutaconyl-CoA.  相似文献   

2.
Cell-free extracts of Pseudomonas sp. strains KB 740 and K 172 both contained high levels of glutaryl-CoA dehydrogenase when grown anaerobically on benzoate or other aromatic compounds and with nitrate as electron acceptor. These aromatic compounds have in common benzoyl-CoA as the central aromatic intermediate of anerobic metabolism. The enzymatic activity was almost absent in cells grown aerobically on benzoate regardless whether nitrate was present. Glutaryl-CoA dehydrogenase activity was also detected in cell-free extracts of Rhodopseudomonas, Rhodomicrobium and Rhodocyclus after phototrophic growth on benzoate. Parallel to the induction of glutaryl-CoA dehydrogenase as measured with ferricenium ion as electron acceptor, an about equally high glutaconyl-CoA decarboxylase activity was detected in cell-free extracts. The latter activity was measured with the NAD-dependent assay, as described for the biotin-containing sodium ion pump glutaconyl-CoA decarboxylase from glutamate fermenting bacteria. Glutaryl-CoA dehydrogenase was purified to homogeneity from both Pseudomonas strains. The enzymes catalyse the decarboxylation of glutaconyl-CoA at about the same rate as the oxidative decarboxylation of glutaryl-CoA. The green enzymes are homotetramers (m=170 kDa) and contain 1 mol FAD per subunit. No inhibition was observed with avidin indicating the absence of biotin. The N-terminal sequences of the enzymes from both strains are similar (65%).  相似文献   

3.
The protonation of crotonyl-CoA dienolate following decarboxylation of glutaconyl-CoA by glutaryl-CoA dehydrogenase was investigated. Although it is generally held that the active sites of acyl-CoA dehydrogenases are desolvated when substrate binds, recent evidence has established that water has access to the active site in these binary complexes of glutaryl-CoA dehydrogenase. The present investigation shows that the dehydrogenase catalyzes (a) a rapid exchange of C-4 methyl protons of crotonyl-CoA with bulk solvent and (b) protonation of crotonyl-CoA dienolate by solvent-derived protons under single turnover conditions. Both of the reactions require the catalytic base, Glu370. These findings indicate that decarboxylation proceeds via a dienolate intermediate. The involvement of water in catalysis by glutaryl-CoA dehydrogenase was previously unrecognized and is in conflict with a classically held intramolecular 1,3-prototropic shift for protonation of crotonyl-CoA dienolate.  相似文献   

4.
The steric course of the decarboxylation of glutaconyl-CoA to crotonyl-CoA, catalysed by the biotin-dependent sodium pump glutaconyl-CoA decarboxylase from Acidaminococcus fermentans, was elucidated using the sequence: chiral acetate----citrate----glutamate----glutaconyl-CoA----crotonyl-CoA ----chiral acetate. Since glutaconyl-CoA or glutaconate labeled at C-4 was subjected to rapid chemical or enzymatic exchanges, glutamate was fermented to acetate by growing cells of A. fermentans. The analysis of the final chiral acetates gave following deviations from 50% in the fumarase exchange: + 13.8% starting with (R)-acetate and - 13.9% starting with (S)-acetate. The results demonstrated a retention of configuration during the decarboxylation. Thus glutaconyl-CoA decarboxylase adds to the list of biotin enzymes in which exclusive retention of configuration was observed. Glutaconate CoA-transferase from A. fermentans catalysed a 3H exchange of [2,4,4-3H]glutaconate with water when acetyl-CoA was present. At low concentration of acetyl-CoA (20 microM) the exchange ceased after exactly one atom 3H was released into the water, at high concentrations (1 mM) the exchange proceeded further. The apparent Km of acetyl-CoA in the exchange (1.1 microM) was 150 times smaller than that of the complete CoA transfer. It was concluded that either a mixed anhydride, between a carboxyl group of the enzyme and [2,4,4-3H]glutaconate, or enzyme-bound glutaconyl-CoA was the exchanging species.  相似文献   

5.
Fu Z  Wang M  Paschke R  Rao KS  Frerman FE  Kim JJ 《Biochemistry》2004,43(30):9674-9684
Acyl-CoA dehydrogenases (ACDs) are a family of flavoenzymes that metabolize fatty acids and some amino acids. Of nine known ACDs, glutaryl-CoA dehydrogenase (GCD) is unique: in addition to the alpha,beta-dehydrogenation reaction, common to all ACDs, GCD catalyzes decarboxylation of glutaryl-CoA to produce CO(2) and crotonyl-CoA. Crystal structures of GCD and its complex with 4-nitrobutyryl-CoA have been determined to 2.1 and 2.6 A, respectively. The overall polypeptide folds are the same and similar to the structures of other family members. The active site of the unliganded structure is filled with water molecules that are displaced when enzyme binds the substrate. The structure strongly suggests that the mechanism of dehydrogenation is the same as in other ACDs. The substrate binds at the re side of the FAD ring. Glu370 abstracts the C2 pro-R proton, which is acidified by the polarization of the thiolester carbonyl oxygen through hydrogen bonding to the 2'-OH of FAD and the amide nitrogen of Glu370. The C3 pro-R proton is transferred to the N(5) atom of FAD. The structures indicate a plausible mechanism for the decarboxylation reaction. The carbonyl polarization initiates decarboxylation, and Arg94 stabilizes the transient crotonyl-CoA anion. Protonation of the crotonyl-CoA anion occurs by a 1,3-prototropic shift catalyzed by the conjugated acid of the general base, Glu370. A tight hydrogen-bonding network involving gamma-carboxylate of the enzyme-bound glutaconyl-CoA, with Tyr369, Glu87, Arg94, Ser95, and Thr170, optimizes orientation of the gamma-carboxylate for decarboxylation. Some pathogenic mutations are explained by the structure. The mutations affect protein folding, stability, and/or substrate binding, resulting in inefficient/inactive enzyme.  相似文献   

6.
We studied the activated oxygen metabolism of peroxisomes in naturally and dark-induced senescent leaves of pea (Pisum sativum L.). Peroxisomes were purified from three different types of senescent leaves and the activities of different peroxisomal and glyoxysomal enzymes were measured. The activities of the O2-- and H2O2-producing enzymes were enhanced by natural senescence. Senescence also produced an increase in the generation of active oxygen species (O2- and H2O2) in leaf peroxisomes and in the activities of two glyoxylate-cycle marker enzymes. A new fraction of peroxisomes was detected at an advanced stage of dark-induced senescence. Electron microscopy revealed that this new peroxisomal fraction varied in size and electron density. During senescence, the constitutive Mn-superoxide dismutase (SOD) activity of peroxisomes increased and two new CuZn-SODs were induced, one of which cross-reacted with an antibody against glyoxysomal CuZn- SOD. This fact and the presence of glyoxylate-cycle enzymes support the idea that foliar senescence is associated with the transition of peroxisomes into glyoxysomes. Our results indicate that natural senescence causes the same changes in peroxisome-activated oxygen metabolism as dark-induced senescence, and reinforce the hypothesis of an effective role of peroxisomes and their activated oxygen metabolism in this stage of the life cycle.  相似文献   

7.
Glutaryl-coenzyme A (CoA) dehydrogenases (GDHs) are acyl-CoA dehydrogenases, which usually dehydrogenate and decarboxylate the substrate to crotonyl-CoA. In some anaerobic bacteria, non-decarboxylating GDHs exist that release glutaconyl-CoA (2,3-dehydroglutaryl-CoA) without decarboxylation. The differing mechanisms of decarboxylating and non-decarboxylating GDHs were investigated by site-directed mutagenesis of the gene coding for the crotonyl-CoA-forming GDH from Geobacter metallireducens. Exchange of single amino acids involved in substrate carboxylate binding impaired the decarboxylation step, resulting in relative glutaconyl-CoA:crotonyl-CoA formation rates of 1:1 (S97A) or 13:1 (Y370A). The total amount of glutaconyl-CoA formed was maximal in the Y370V+S97A double mutant. The results obtained indicate that an invariant deprotonated Tyr plays a crucial role for optimizing the leaving group potential of CO(2) in decarboxylating GDHs.  相似文献   

8.
W Buckel  R Semmler 《FEBS letters》1982,148(1):35-38
The decarboxylation of glutaconyl-CoA to crotonyl-CoA in the anaerobic bacterium Acidaminococcus fermentans is catalysed by a membrane-bound, biotin-dependent enzyme which requires Na+ for activity. Inverted vesicles from A. fermentans accumulated Na+ only if glutaconyl-CoA was decarboxylated. The Na+ uptake was inhibited by avidin but not by the avidin biotin complex. Detergents and ionophores such as monensin also prevented the Na+ transport. The results indicate that the enzyme is able to convert the free energy of decarboxylation (delta Go' approximately equal to -30 kJ/mol) into a Na+ gradient.  相似文献   

9.
Wendt KS  Schall I  Huber R  Buckel W  Jacob U 《The EMBO journal》2003,22(14):3493-3502
Glutaconyl-CoA decarboxylase is a biotin-dependent ion pump whereby the free energy of the glutaconyl-CoA decarboxylation to crotonyl-CoA drives the electrogenic transport of sodium ions from the cytoplasm into the periplasm. Here we present the crystal structure of the decarboxylase subunit (Gcdalpha) from Acidaminococcus fermentans and its complex with glutaconyl-CoA. The active sites of the dimeric Gcdalpha lie at the two interfaces between the mono mers, whereas the N-terminal domain provides the glutaconyl-CoA-binding site and the C-terminal domain binds the biotinyllysine moiety. The Gcdalpha catalyses the transfer of carbon dioxide from glutaconyl-CoA to a biotin carrier (Gcdgamma) that subsequently is decarboxylated by the carboxybiotin decarboxylation site within the actual Na(+) pump (Gcdbeta). The analysis of the active site lead to a novel mechanism for the biotin-dependent carboxy transfer whereby biotin acts as general acid. Furthermore, we propose a holoenzyme assembly in which the water-filled central channel of the Gcdalpha dimer lies co-axial with the ion channel (Gcdbeta). The central channel is blocked by arginines against passage of sodium ions which might enter the central channel through two side channels.  相似文献   

10.
To clarify the significance of catalase in peroxisomes, we have examined the effect of aminotriazole treatment of rats on the activity of beta-hydroxybutyryl-CoA dehydrogenase in liver peroxisomes. When the effect of H2O2 on the dehydrogenase activity was examined using an extract of liver peroxisomes from aminotriazole-treated rats, the acetoacetyl-CoA-dependent oxidation of NADH was found to increase considerably on the addition of dilute H2O2. Such an effect of H2O2 was not seen on the beta-hydroxybutyryl-CoA-dependent reduction of NAD nor with extracts from untreated animals. We then noticed that similar NADH oxidation was caused non-enzymatically by a mixture of acetoacetyl-CoA and H2O2. The oxidation was dependent on both acetoacetyl-CoA and H2O2, and was blocked by scavengers of oxyradicals such as ascorbate and ethanol. Degradation products formed during the reaction of acetoacetyl-CoA with H2O2 had no NADH oxidizing activity, indicating that effective oxidant(s) were generated during the reaction of H2O2 with acetoacetyl-CoA. No other fatty acyl-CoA so far examined nor acetoacetate could replace acetoacetyl-CoA in this reaction. Therefore, if H2O2 were to be accumulated in peroxisomes, it would decrease both NADH and acetoacetyl-CoA, thus affecting the fatty acyl-CoA beta-oxidation system. These results, together with our previous finding that peroxisomal thiolase was significantly inactivated by H2O2 [Hashimoto, F. & Hayashi, H. (1987) Biochim. Biophys. Acta 921, 142-150] suggest that the role of catalase in peroxisomes is at least in part to protect the fatty acyl-CoA beta-oxidation system from the deleterious action of H2O2.  相似文献   

11.
A spectroscopic method employing pulsed dye laser instrumentation is described for the determination of the 1H2O/2H2O composition of aqueous solutions by the measurement of reciprocal excited state lifetimes of EuEDTA-. The reciprocal lifetimes, gamma-1, of the 1H2O/2H2O mixtures increase linearly with the mole fraction of 1H2O. For EuEDTA- the relationship between gamma-1 and the mole fraction, chi H, of 1H2O in 1H2O/2H2O mixtures is expressed by the equation chi H = 0.37 gamma-1-0.152, with a sensitivity in chi H of +/- 0.02. The reciprocal lifetimes are independent of pH in the range 5.1 to 10.5, changes in ionic strength, and the type of buffer used in EuEDTA- containing solutions.  相似文献   

12.
Glutaryl-CoA dehydrogenase, a multifunctional enzyme responsible for dehydrogenation and decarboxylation of glutaryl-CoA to crotonyl-CoA, has been purified 1,680-fold from porcine liver mitochondria. The purified porcine enzyme has a subunit molecular weight of 47,800 and a native molecular weight of 190,500. Porcine glutaryl-CoA dehydrogenase catalyzed the conversion of [1,5-14C]glutaryl-CoA to [14C] crotonyl-CoA and 14CO2 in a 1:1:1 ratio. The porcine enzyme has Km values for electron transfer flavoprotein and glutaryl-CoA of 1.1 and 3.3 microM, respectively, and turnover numbers of 860 mol of electron transfer flavoprotein/min/mol of glutaryl-CoA dehydrogenase and 327 mol of glutaryl-CoA/min/mol of glutaryl-CoA dehydrogenase. Human glutaryl-CoA dehydrogenase has been purified 1,278-fold from human liver mitochondria. The purified human enzyme has a subunit molecular weight of 58,800 and a native molecular weight of 256,000. Human glutaryl-CoA dehydrogenase showed a reaction of only partial identity when compared to porcine glutaryl-CoA dehydrogenase by Ouchterlony double immunodiffusion analysis using antiserum raised against and monospecific for porcine glutaryl-CoA dehydrogenase.  相似文献   

13.
Glutaconyl-CoA decarboxylase from Acidaminococcus fermentans was inactivated by incubation with n-alkanols at 37 degrees C. The concentration of the alcohol required for complete inactivation decreased with increasing chain length; e.g. 2 M ethanol was as potent as 2 mM hexanol or 0.5 mM decanol. The data indicate a binding of the alcohol to the enzyme with an energy of about 4 kJ/methylene group. Sodium ions prevented the inactivation (50% at 30 mM NaCl). K+, NH4+, Cs+ and Mg2+ had no influence, whereas Li+ was ten times less effective than Na+. The enzyme was cleaved during the inactivation into a soluble part, consisting of the alpha (Mr 120,000) and beta polypeptide chains (60,000), whereas the hydrophobic gamma chain (30,000) precipitated. The soluble part catalysed the sodium-ion-independent but avidin-sensitive glutaconyl-CoA/crotonyl-CoA exchange as measured with the substrates [3-3H]crotonyl-CoA and unlabelled glutaconate and with glutaconate CoA-transferase as auxiliary enzyme. In the presence of free biotin or its methyl ester the soluble part catalysed the formation of crotonyl-CoA from glutaconyl-CoA (apparent Km for biotin 40 mM, Vmax 1% of the native decarboxylation reaction). This apparent reactivation was most likely caused by the carboxylation of free biotin. Based on these and other observations the following functions may be assigned to the different polypeptide chains of glutaconyl-CoA decarboxylase: biotin carrier (alpha), carboxytransferase (beta) and carboxylase, the actual sodium pump (gamma).  相似文献   

14.
Dwyer TM  Rao KS  Goodman SI  Frerman FE 《Biochemistry》2000,39(37):11488-11499
Glutaryl-CoA dehydrogenase catalyzes the oxidation of glutaryl-CoA to crotonyl-CoA and CO(2) in the mitochondrial degradation of lysine, hydroxylysine, and tryptophan. We have characterized the human enzyme that was expressed in Escherichia coli. Anaerobic reduction of the enzyme with sodium dithionite or substrate yields no detectable semiquinone; however, like other acyl-CoA dehydrogenases, the human enzyme stabilizes an anionic semiquinone upon reduction of the complex between the enzyme and 2,3-enoyl-CoA product. The flavin potential of the free enzyme determined by the xanthine-xanthine oxidase method is -0.132 V at pH 7.0, slightly more negative than that of related flavoprotein dehydrogenases. A single equivalent of substrate reduces 26% of the dehydrogenase flavin, suggesting that the redox equilibrium on the enzyme between substrate and product and oxidized and reduced flavin is not as favorable as that observed with other acyl-CoA dehydrogenases. This equilibrium is, however, similar to that observed in isovaleryl-CoA dehydrogenase. Comparison of steady-state kinetic constants of glutaryl-CoA dehydrogenase with glutaryl-CoA and the alternative substrates, pentanoyl-CoA and hexanoyl-CoA, suggests that the gamma-carboxyl group of glutaryl-CoA stabilizes the enzyme-substrate complex by at least 5.7 kJ/mol, perhaps by interaction with Arg94 or Ser98. Glu370 is positioned to function as the catalytic base, and previous studies indicate that the conjugate acid of Glu370 also protonates the transient crotonyl-CoA anion following decarboxylation [Gomes, B., Fendrich, G. , and Abeles, R. H. (1981) Biochemistry 20, 3154-3160]. Glu370Asp and Glu370Gln mutants of glutaryl-CoA dehydrogenase exhibit 7% and 0. 04% residual activity, respectively, with human electron-transfer flavoprotein; these mutations do not grossly affect the flavin redox potentials of the mutant enzymes. The reduced catalytic activities of these mutants can be attributed to reduced extent and rate of substrate deprotonation based on experiments with the nonoxidizable substrate analogue, 3-thiaglutaryl-CoA, and kinetic experiments. Determination of these fundamental properties of the human enzyme will serve as the basis for future studies of the decarboxylation reaction which is unique among the acyl-CoA dehydrogenases.  相似文献   

15.
B Gomes  G Fendrich  R H Abeles 《Biochemistry》1981,20(6):1481-1490
Glutaryl-CoA dehydrogenase, a flavoprotein, catalyzes the reaction -OOCCH3CH2--CH2COSR (FAD leads to FADH2) leads to CH3CH = CHCOSR + CO2 (SR = CoA or pantetheine). With the isolated enzyme, a dye serves as the final electron acceptor. The enzyme from Pseudomonas fluorescens (ATCC 11250) has been purified to homogeneity. It was established with appropriate isotopic substitutions that the proton which is added to the gamma position of the product, subsequent to decarboxylation, is not derived from the solvent but is derived from the alpha position of the substrate. Under conditions where no net conversion of substrate occurs, i.e., in the absence of electron acceptor, the enzyme catalyzes the exchange of the beta hydrogen of the substrate with solvent protons. Butyryl-CoA dehydrogenase (M. elsedenii), which catalyzes an analogous reaction, catalyzes the exchange of both the alpha and beta hydrogens with solvent protons in the absence of electron acceptor. Glutaryl-CoA dehydrogenase and butyryl-CoA dehydrogenase are irreversibly inactivated by the substrate analogues 3-butynoylpantetheine and 3-pentynoylpantetheine. These inactivators do not form an adduct with the flavin and probably react with a nucleophile at the active site. Upon inactivation, the spectrum of the enzyme-bound flavin is essentially unchanged, and the flavin can be reduced by Na2S2O4. We suggest that inactivation involves intermediate allene formation. We proposed that these results support an oxidation mechanism for glutaryl-CoA dehydrogenase and butyryl-CoA dehydrogenase which is initiated by proton abstraction. With glutaryl-CoA dehydrogenase, the base, which abstracts the substrate alpha proton, is shielded from the solvent and is then used to protonate the carbanion (CH2--CH--CHCOSCoA) formed after oxidation and decarboxylation.  相似文献   

16.
The exchange of oxygen atoms between acetate, glutaryl-CoA, and the catalytic glutamate residue in glutaconate CoA-transferase from Acidaminococcus fermentans was analyzed using [(18)O(2)]acetate together with matrix-assisted laser desorption/ionization time of flight mass spectrometry of an appropriate undecapeptide. The exchange reaction was shown to be site-specific, reversible, and required both glutaryl-CoA and [(18)O(2)]acetate. The observed exchange is in agreement with the formation of a mixed anhydride intermediate between the enzyme and acetate. In contrast, with a mutant enzyme, which was converted to a thiol ester hydrolyase by replacement of the catalytic glutamate residue by aspartate, no (18)O uptake from H(2)(18)O into the carboxylate was detectable. This result is in accord with a mechanism in which the carboxylate of aspartate acts as a general base in activating a water molecule for hydrolysis of the thiol ester intermediate. This mechanism is further supported by the finding of a significant hydrolyase activity of the wild-type enzyme using acetyl-CoA as substrate, whereas glutaryl-CoA is not hydrolyzed. The small acetate molecule in the substrate binding pocket may activate a water molecule for hydrolysis of the nearby enzyme-CoA thiol ester.  相似文献   

17.
The existence of a relationship between clofibrate-induced peroxisome proliferation and oxidative stress mediated by activated oxygen species was studied in intact peroxisomes purified from Pisum sativum L. plants. Incubation of leaves with 1 mM clofibrate produced a remarkable increase in the peroxisomal activity of acyl-CoA oxidase and, to a lesser extent, of xanthine oxidase, whereas there was a nearly complete loss of catalase activity and a decrease in Mn-superoxide dismutase. Ultrastructural studies of intact leaves showed that clofibrate induced a five- and twofold proliferation of the peroxisomal and mitochondrial populations, respectively, in comparison with those in control leaves. Prolonged incubation with clofibrate produced considerable alterations in the ultrastructure of cells. In peroxisomal membranes, the NADH-induced generation of O2- radicals, as well as the lipid peroxidation of membranes, increased as a result of treatment of plants with clofibrate. In intact peroxisomes treated with this hypolipidemic drug, the H2O2 concentration was higher than in peroxisomes from control plants. These results demonstrate that clofibrate stimulates the production of activated oxygen species (O2- and H2O2) inside peroxisomes, as well as the lipid peroxidation of peroxisomal membranes. This effect is concomitant with a decrease of catalase and Mn-SOD activities, the main peroxisomal enzymatic defenses against H2O2 and O2-, and indicates that in the toxicity of clofibrate, at the level of peroxisomes, an oxidative stress mechanism mediated by activated oxygen species is involved.  相似文献   

18.
In plant cells, as in most eukaryotic organisms, peroxisomes are probably the major sites of intracellular H2O2 production, as a result of their essentially oxidative type of metabolism. Like mitochondria and chloroplasts, peroxisomes also produce superoxide radicals (O2*-) and there are, at least, two sites of superoxide generation: one in the organelle matrix, the generating system being xanthine oxidase, and another site in the peroxisomal membranes dependent on NAD(P)H. In peroxisomal membranes, three integral polypeptides (PMPs) with molecular masses of 18, 29, and 32 kDa have been shown to generate O2*- radicals. Besides catalase, several antioxidative systems have been demonstrated in plant peroxisomes, including different superoxide dismutases, the four enzymes of the ascorbate-glutathione cycle plus ascorbate and glutathione, and three NADP-dependent dehydrogenases. A CuZn-SOD and two Mn-SODs have been purified and characterized from different types of plant peroxisomes. The presence of the enzyme nitric oxide synthase (NOS) and its reaction product, nitric oxide (NO*), has been recently demonstrated in plant peroxisomes. Different experimental evidence has suggested that peroxisomes have a ROS-mediated cellular function in leaf senescence and in stress situations induced by xenobiotics and heavy metals. Peroxisomes could also have a role in plant cells as a source of signal molecules like NO*, O2*- radicals, H2O2, and possibly S-nitrosoglutathione (GSNO). It seems reasonable to think that a signal molecule-producing function similar to that postulated for plant peroxisomes could also be performed by human, animal and yeast peroxisomes, where research on oxy radicals, antioxidants and nitric oxide is less advanced than in plant peroxisomes.  相似文献   

19.
通过解析葡萄糖有氧氧化过程中有机酸脱羧生成CO2的全部O原子来源,以及呼吸链一氧化磷酸化生成ATP的葡萄糖以外H原子的来源,明晰了葡萄糖有氧氧化过程中直接或者间接加H2O的特殊意义:H2O的H原子进入呼吸链一氧化磷酸化释放能量生成ATP;O原子结合到中间产物的C原子上形成羧基一COOH,以有机酸脱酸形式生成CO2释放出来。  相似文献   

20.
To determine the steric course of the reaction of bacterial ornithine decarboxylase [EC 4.1.1.17], we have carried out the decarboxylation of L-ornithine in 2H2O and that of DL-[2-2H]ornithine in H2O, and obtained putrescine bearing a single deuterium atom in the C-1 position. The stereochemistry of [1-2H]putrescine was established by conversion to 1-(2-pyrrolidinyl)-2-propanone with acetoacetate and the pro-S hydrogen-specific diamine oxidase from pea seedlings. Analysis of deuterium content by gas chromatography-mass spectrometry showed that the deuterium label was fully retained during the conversion of [1-2H]putrescine produced by the decarboxylation of L-ornithine in 2H2O to 1-(2-pyrrolidinyl)-2-propanone, in contrast with the considerable loss of label from [1-2H]putrescine which was produced by the decarboxylation of DL-[2-2H]ornithine in H2O. The extent of loss of the deuterium label was in good agreement with the estimated value based on the isotope effect in the diamine oxidase reaction. These results indicate that the introduced deuterium (or hydrogen) is in the pro-R position at C-1 of putrescine, and consequently the ornithine decarboxylase reaction proceeds with retention of configuration.  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