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.
Xiong H  Stanley BA  Pegg AE 《Biochemistry》1999,38(8):2462-2470
S-Adenosylmethionine decarboxylase is a pyruvate-dependent enzyme. The enzyme forms a Schiff base with substrate, S-adenosylmethionine, through the pyruvoyl moiety. This facilitates the release of CO2 from the substrate, which must then be protonated on the alpha carbon in order to permit hydrolysis of the Schiff base to release the product. The catalytic mechanism of human S-adenosylmethionine decarboxylase was investigated via mutagenic and kinetic approaches. The results of enzyme kinetic studies indicated that Cys-82 is a crucial residue for activity and this residue has a basic pKa. Iodoacetic acid inhibited wild-type enzyme activity in a time- and pH-dependent manner but did not affect the already reduced activity of mutant C82A. Reaction of this mutant with iodoacetic acid led to approximately one less mole of reagent being incorporated per mole of enzyme alphabeta dimer than with wild-type S-adenosylmethionine decarboxylase. Both wild-type and C82A mutant S-adenosylmethionine decarboxylases were inactivated by substrate-mediated transamination, but this reaction occurred much more frequently with C82A than with wild-type enzyme. A major proportion of the recombinant C82A mutant protein was in the transaminated form in which the pyruvoyl cofactor is converted into alanine. This suggests that incorrect protonation of the pyruvate, rather than the substrate, occurs much more readily when Cys-82 is altered. On the basis of these results, it was postulated that residue Cys-82 may be the proton donor of the decarboxylation reaction catalyzed by S-adenosylmethionine decarboxylase.  相似文献   

2.
S-Adenosylmethionine decarboxylase is one of a small group of enzymes that use a pyruvoyl residue as a cofactor. Histidine decarboxylase from Lactobacillus 30a, the best studied pyruvoyl-containing enzyme, has an (alpha beta)6 subunit structure with the pyruvoyl moiety linked through an amide bond to the NH2-terminal of the larger alpha subunit (Recsei, P. A., Huynh, Q. K., and Snell, E. E. (1983) Proc. Natl. Acad. Sci. U. S. A. 80, 973-977). To examine potential structural analogies between the two enzymes, we have isolated and partially characterized S-adenosylmethionine decarboxylase. The purified enzyme comprises equimolar amounts of two subunits of Mr = 14,000 and 19,000 (by sodium dodecyl sulfate-polyacrylamide gel electrophoresis) and has a native molecular weight of 136,000 (by gel filtration). Approximately 4 mol of [methyl-3H] adenosylmethionine are incorporated per mol of enzyme (Mr = 136,000) when the enzyme is inactivated with this substrate and NaCNBH3. These data suggest an (alpha beta)4 structure with 1 pyruvoyl residue for each alpha beta pair. The two subunits have been separated by reversed-phase high performance liquid chromatography after reduction and carboxymethylation. The smaller subunit (beta) has a free amino terminus. The amino terminus of the larger subunit (alpha) appears to be blocked by a pyruvoyl group; this subunit can be sequenced only after this group is converted to an alanyl residue by reduction with sodium cyanoborohydride in the presence of ammonium acetate. This work suggests that S-adenosylmethionine decarboxylase is structurally much more similar to histidine decarboxylase than previously thought.  相似文献   

3.
The established pathways from serine to ethanolamine are indirect and involve decarboxylation of phosphatidylserine. Here we show that plants can decarboxylate serine directly. Using a radioassay based on ethanolamine (Etn) formation, pyridoxal 5'-phosphate-dependent l-serine decarboxylase (SDC) activity was readily detected in soluble extracts from leaves of diverse species, including spinach, Arabidopsis, and rapeseed. A putative Arabidopsis SDC cDNA was identified by searching GenBank for sequences homologous to other amino acid decarboxylases and shown by expression in Escherichia coli to encode a soluble protein with SDC activity. This cDNA was further authenticated by complementing the Etn requirement of a yeast psd1 psd2 mutant. In a parallel approach, a cDNA was isolated from a rapeseed library by its ability to complement the Etn requirement of a yeast cho1 mutant and shown by expression in E. coli to specify SDC. The deduced Arabidopsis and rapeseed SDC polypeptides are 90% identical, lack obvious targeting signals, and belong to amino acid decarboxylase group II. Recombinant Arabidopsis SDC was shown to exist as a tetramer and to contain pyridoxal 5'-phosphate. It does not attack d-serine, l-phosphoserine, other l-amino acids, or phosphatidylserine and is not inhibited by Etn, choline, or their phosphoesters. As a soluble, pyridoxal 5'-phosphate enzyme, SDC contrasts sharply with phosphatidylserine decarboxylases, which are membrane proteins that have a pyruvoyl cofactor.  相似文献   

4.
The maximal turnover rate of CO2 hydration catalyzed by the carbonic anhydrases is limited by proton transfer steps from the zinc-bound water to solution, steps that regenerate the catalytically active zinc-bound hydroxide. Catalysis of CO2 hydration by wild-type human carbonic anhydrase III (HCA III) (k(cat) = 2 ms (-1)) is the least efficient among the carbonic anhydrases in its class, in part because it lacks an efficient proton shuttle residue. We have used site-directed mutagenesis to test positions within the active-site cavity of HCA III for their ability to carry out proton transfer by replacing various residues with histidine. Catalysis by wild-type HCA III and these six variants was determined from the initial velocity of hydration of CO2 measured by stopped-flow spectrophotometry and from the exchange of 18O between CO2 and H2O at chemical equilibrium by mass spectrometry. The results show that histidine at three positions (Lys64His, Arg67His and Phe131His) have the capacity to transfer protons during catalysis, enhancing maximal velocity of CO2 hydration and 18O exchange from 4- to 15-fold compared with wild-type HCA III. Histidine residues at the other three positions (Trp5His, Tyr7His, Phe20His) showed no firm evidence for proton transfer. These results are discussed in terms of the stereochemistry of the active-site cavity and possible proton transfer pathways.  相似文献   

5.
S-Adenosylmethionine decarboxylase (AdoMetDC) is a pyruvoyl-dependent enzyme that catalyzes the formation of the aminopropyl group donor in the biosynthesis of the polyamines spermidine and spermine. The enzyme is synthesized as a protein precursor and is activated by an autocatalytic serinolysis reaction that creates the pyruvoyl group. The autoprocessing reaction proceeds via an N --> O acyl rearrangement, generating first an oxyoxazolidine anion intermediate followed by an ester intermediate. A similar strategy is utilized in self-catalyzed protein splicing reactions and in autoproteolytic activation of protein precursors. Mutation of Ser68 to alanine in human AdoMetDC prevents processing by removing the serine side chain necessary for nucleophilic attack at the adjacent carbonyl carbon atom. We have determined the X-ray structure of the S68A mutant and have constructed models of the proenzyme and the oxyoxazolidine intermediate. Formation of the oxyoxazolidine intermediate is promoted by a hydrogen bond from Cys82 and stabilized by a hydrogen bond from Ser229. These observations are consistent with mutagenesis studies, which show that the C82S and C82A mutants process slowly and that the S229A mutant does not process at all. Donation of a proton by His243 to the nitrogen atom of the oxyoxazolidine ring converts the oxyoxazolidine anion to the ester intermediate. The absence of a base to activate the hydroxyl group of Ser68 suggests that strain may play a role in the cleavage reaction. Comparison of AdoMetDC with other self-processing proteins shows no common structural features. Comparison to histidine decarboxylase and aspartate decarboxylase shows that these pyruvoyl-dependent enzymes evolved different catalytic strategies for forming the same cofactor.  相似文献   

6.
Pyruvoyl‐dependent histidine decarboxylases are produced as proenzymes that mature by cleavage followed by formation of the pyruvoyl prosthetic group. The histidine decarboxylation pathway of Streptococcus thermophilus CHCC1524 that consists of the pyruvoyl‐dependent histidine decarboxylase HdcA and the histidine/histamine exchanger HdcP was functionally expressed in Lactococcus lactis. The operon encoding the pathway contains in addition to the hdcA and hdcP genes a third gene hdcB. Expression of different combinations of the genes in L. lactis and Escherichia coli followed by analysis of the protein products demonstrated the involvement of HdcB in the cleavage of the HdcA proenzyme. The HdcA proenzyme and HdcB protein were purified to homogeneity and cleavage and activation of the histidine decarboxylase activity was demonstrated in vitro. Substoichiometric amounts of HdcB were required to cleave HdcA showing that HdcB functions as an enzyme. In agreement, expression levels of HdcB in the cells were low relative to those of HdcA. The turnover number of HdcB in vitro was extremely low (0.05 min?1) which was due to a very slow association/dissociation of the enzyme/substrate complex. In fact, HdcB was shown to co‐purify both with the HdcA S82A mutant that mimics the proenzyme and with the mature HdcA complex.  相似文献   

7.
Benzoylformate decarboxylase (BFD) enhances the rate of decarboxylation of its key intermediate compared to the nonenzymic reaction by a factor of about 106. It has been proposed that desolvation into a hydrophobic environment will lower the reaction barrier in TDP-dependent decarboxylases. The competition of thiamin thiazolone diphosphate (TTDP) with the cofactor thiamin diphosphate (TDP) provides a dynamic indication of the relative hydrophobicity of the cofactor binding site. BFD binds the more polar TDP tightly in the presence of excess TTDP. Therefore, desolvation would not be likely to occur during catalysis. Unlike TDP enzymes that have electron acceptors as substrates, decarboxylases require protonation to produce the precursor to the aldehyde product. A mechanism involving an associated acid that traps the carbanion generated upon C–C bond breaking will permit diffusional separation of carbon dioxide and generate the appropriate precursor to the product aldehdye. This would also account for avoidance of a competitive reaction. Hasson’s detailed structure of BFD shows a highly polar active site with histidines in the vicinity of the substrate. Reports of a reduction of kcat to near the nonenzymic rate without a large effect on Km upon specific replacement of these histidines with alanine fit this alternative. In TDP enzymes involving oxidation or condensation, an electrophilic substrate or second cofactor will be bound (and no proton will be required). This will acquire the electron density of the carbanion itself. In such cases, protonated side chains are not functional while hydrophobic environments would promote the internal transfer.  相似文献   

8.
The short-lived enzyme S-adenosylmethionine decarboxylase uses a covalently bound pyruvoyl cofactor to catalyze the formation of decarboxylated S-adenosylmethionine, which then donates an aminopropyl group for polyamine biosynthesis. Here we demonstrate that S-adenosylmethionine decarboxylase is ubiquitinated and degraded by the 26 S proteasome in vivo, a process that is accelerated by inactivation of S-adenosylmethionine decarboxylase by substrate-mediated transamination of its pyruvoyl cofactor. Proteasome inhibition in COS-7 cells prevents the degradation of S-adenosylmethionine decarboxylase antigen; however, even brief inhibition of the 26 S proteasome caused substantial losses of S-adenosylmethionine decarboxylase activity despite accumulation of S-adenosylmethionine decarboxylase antigen. Levels of the enzyme's substrate (S-adenosylmethionine) increased rapidly after 26 S proteasome inhibition, and this increase in substrate level is consistent with the observed loss of activity arising from an increased rate of inactivation by substrate-mediated transamination. Evidence is also presented that this substrate-mediated transamination accelerates normal degradation of S-adenosylmethionine decarboxylase, as the rate of degradation of the enzyme was increased in the presence of AbeAdo (5'-([(Z)-4-amino-2-butenyl]methylamino]-5'-deoxyadenosine) (a substrate analogue that transaminates the enzyme); conversely, when the intracellular substrate level was reduced by methionine deprivation, the rate of degradation of the enzyme was decreased. Ubiquitination of S-adenosylmethionine decarboxylase is demonstrated by isolation of His-tagged AdoMetDC (S-adenosylmethionine decarboxylase) from COS-7 cells co-transfected with hemagglutinin-tagged ubiquitin and showing bands that were immunoreactive to both anti-AdoMetDC antibody and anti-hemagglutinin antibody. This is the first study to demonstrate that AdoMetDC is ubiquitinated and degraded by the 26 S proteasome, and substrate-mediated acceleration of degradation is a unique finding.  相似文献   

9.
The review is concerned with three Na(+)-dependent biotin-containing decarboxylases, which catalyse the substitution of CO(2) by H(+) with retention of configuration (DeltaG degrees '=-30 kJ/mol): oxaloacetate decarboxylase from enterobacteria, methylmalonyl-CoA decarboxylase from Veillonella parvula and Propiogenium modestum, and glutaconyl-CoA decarboxylase from Acidaminococcus fermentans. The enzymes represent complexes of four functional domains or subunits, a carboxytransferase, a mobile alanine- and proline-rich biotin carrier, a 9-11 membrane-spanning helix-containing Na(+)-dependent carboxybiotin decarboxylase and a membrane anchor. In the first catalytic step the carboxyl group of the substrate is converted to a kinetically activated carboxylate in N-carboxybiotin. After swing-over to the decarboxylase, an electrochemical Na(+) gradient is generated; the free energy of the decarboxylation is used to translocate 1-2 Na(+) from the inside to the outside, whereas the proton comes from the outside. At high [Na(+)], however, the decarboxylases appear to catalyse a mere Na(+)/Na(+) exchange. This finding has implications for the life of P. modestum in sea water, which relies on the synthesis of ATP via Delta(mu)Na(+) generated by decarboxylation. In many sequenced genomes from Bacteria and Archaea homologues of the carboxybiotin decarboxylase from A. fermentans with up to 80% sequence identity have been detected.  相似文献   

10.
11.
Glutamate decarboxylase is a pyridoxal 5'-phosphate (PLP)-dependent enzyme, belonging to the subset of PLP-dependent decarboxylases classified as group II. Site-directed mutagenesis of Escherichia coli glutamate decarboxylase, combined with analysis of the crystal structure, shows that a histidine residue buried in the protein core is critical for correct folding. This histidine is strictly conserved in the PF00282 PFAM family, which includes the group II decarboxylases. A similar role is proposed for residue Ser269, also highly conserved in this group of enzymes, as it provides one of the interactions stabilising His241.  相似文献   

12.
Mammalian histidine decarboxylases have not been characterized well owing to their low amounts in tissues and instability. We describe here the first spectroscopic characterization of a mammalian histidine decarboxylase, i.e. a recombinant version of the rat enzyme purified from transformed Escherichia coli cultures, with similar kinetic constants to those reported for mammalian histidine decarboxylases purified from native sources. We analyzed the absorption, fluorescence and circular dichroism spectra of the enzyme and its complexes with the substrate and substrate analogues. The pyridoxal-5'-phosphate-enzyme internal Schiff base is mainly in an enolimine tautomeric form, suggesting an apolar environment around the coenzyme. Michaelis complex formation leads to a polarized, ketoenamine form of the Schiff base. After transaldimination, the coenzyme-substrate Schiff base exists mainly as an unprotonated aldimine, like that observed for dopa decarboxylase. However, the coenzyme-substrate Schiff base suffers greater torsion than that observed in other L-amino acid decarboxylases, which may explain the relatively low catalytic efficiency of this enzyme. The active center is more resistant to the formation of substituted aldamines than the prokaryotic homologous enzyme and other L-amino acid decarboxylases. Characterization of the similarities and differences of mammalian histidine decarboxylase with respect to other homologous enzymes would open new perspectives for the development of new and more specific inhibitors with pharmacological potential.  相似文献   

13.
Various bacteria are able to grow aerobically or anaerobically on malonate as sole source of carbon and energy. Independent of the mechanism for energy conservation, the decarboxylation of malonate is the key reaction in the decomposition of this compound. To achieve malonate decarboxylation under physiological conditions, the substrate must be converted into an activated (thioester) derivative. We report here on the malonate decarboxylases of Malonomonas rubra and Klebsiella pneumoniae. These enzymes perform an interesting substrate activation mechanism by generating a malonyl thioester with the enzyme. Formation of the malonyl-S-enzyme involves an 'activation module' that comprises the acetylation of a specific thiol group of an acyl carrier protein (ACP) and the transfer of the ACP moiety to malonate, yielding malonyl-S-ACP and acetate. The malonyl-S-ACP is subsequently decarboxylated with regeneration of the acetyl-ACP. The malonate activation mechanism is related to the activation of citrate by citrate lyase. The relationship extends to the identical 2'-(5'-phosphoribosyl)-3'-dephospho-CoA thiol cofactor that is bound covalently to the corresponding ACP subunit. In Klebsiella pneumoniae, malonate is decarboxylated by a water-soluble enzyme complex. In the anaerobic bacterium Malonomonas rubra, malonate decarboxylation is catalysed by a set of water-soluble as well as membrane-bound enzymes that function together in converting the free energy of the decarboxylation reaction into delta muNa+. Therefore, this malonate decarboxylase includes a biotin carrier protein that accepts the CO2 moiety from malonyl-S-ACP and delivers it to a membrane-bound decarboxylase acting as a Na+ pump. Genes encoding the individual protein components that perform the decarboxylation of malonate in K. pneumoniae or M. rubra have been identified within the mdc and mad gene clusters respectively. The function of most of the derived proteins could be envisaged from sequence similarities with proteins of known functions. The genetic evidence firmly supports the idea that malonate decarboxylation is carried out by the two different decarboxylases, as deduced from the biochemical studies of the enzymes.  相似文献   

14.
The mechanism and stereochemistry in connection with enzymatic conversion of cholesta-4,6-dien-3-one into cholestanol was studied. Rat and mouse liver microsomes are able to catalyze NADPH-dependent sequential saturation of the two double bonds. Evidence was obtained that the saturation of the delta 6-double bond includes transfer of a hydride ion from the B-side of the cofactor to the 7-position of the steroid (mainly 7 beta-position), followed by addition of a proton to the 6 alpha-position (mainly trans addition). The saturation of the delta 4-double bond includes transfer of a hydride ion from the B-side of the cofactor to the 5 alpha-position of the steroid followed by addition of a proton to the 4 beta-position (trans addition). The reduction of the 3-oxo group was found to involve transfer of a hydride ion from the B-side of the cofactor NADPH to the 3 alpha-position of the steroid. The results are in accord with the contention that the enzymatic saturation of the two double bonds involves a polarization of the 3-oxo group making C-7 electrophilic and C-6 nucleophilic in connection with the saturation of the delta 6-double bond and C-5 electrophilic and C-4 nucleophilic in connection with the saturation of the delta 4-double bond.  相似文献   

15.
The ubiquinol oxidase cytochrome bo3 from Escherichia coli is one of the respiratory heme-copper oxidases which catalyze the reduction of O2 to water linked to translocation of protons across the bacterial or mitochondrial membrane. We have studied the structure of the CuB site in the binuclear heme-copper center of O2 reduction by EXAFS spectroscopy in the fully reduced state of this enzyme, as well as in the reduced CO-liganded states where CO is bound either to the heme iron or to CuB. We find that, in the reduced enzyme, CuB is coordinated by one weakly bound and two strongly bound histidine imidazoles at Cu-N distances of 2.10 and 1.92 A, respectively, and that an additional feature at 2.54 A is due to a highly ordered water molecule that might be weakly associated with the copper. Unexpectedly, the binding of CO to heme iron is found to result in a major conformational change at CuB, which now binds only two equidistant histidine imidazoles at 1.95 A and a chloride ion at 2. 25 A, with elimination of the water molecule and one of the histidines. Attempts to remove the chloride from the enzyme by extensive dialysis did not change this finding, nor did substitution of chloride with bromide. Photolysis of CO bound to the heme iron is known to cause the CO to bind to CuB in a very fast reaction and to remain bound to CuB at low temperatures. In this state, we indeed find the CO to be bound to CuB at a Cu-C distance of 1.85 A, with chloride still bound at 2.25 A and the two histidine imidazoles at a Cu-N distance of 2.01 A. These results suggest that reduction of the binuclear site weakens the bond between CuB and one of its three histidine imidazole ligands, and that binding of CO to the reduced binuclear site causes a major structural change in CuB in which one histidine ligand is lost and replaced by a chloride ion. Whether chloride is a cofactor in this enzyme is discussed.  相似文献   

16.
The two cysteinyl residues present in histidine decarboxylase from Lactobacillus 30a differ greatly in reactivity. One (class 1) reacts readily in the native state with dithiobis-(2-nitrobenzoate) with complete loss of enzyme activity; the other (class 2) reacts only after denaturation of the enzyme (Lane, R. S., and Snell, E. E. (1976) Biochemistry 15, 4175-4179). These differences in reactivity permitted use of covalent (disulfide) chromatography to isolate separate peptides that contain these two residues. Sequence analysis showed that the class 1 cysteinyl residue is at position 147 in a hydrophilic portion of the alpha chain (Huynh, Q. K., Recsei, P. A., Vaaler, G. L., and Snell, E. E. (1984) J. Biol. Chem. 259, 2833-2839), while the class 2 cysteinyl residue is present at position 71, adjacent to a hydrophobic portion of the same chain. Cysteinyl peptides identical with or homologous to the class 2 cysteinyl peptide of the Lactobacillus 30a enzyme were isolated from the alpha subunits of histidine decarboxylases from Lactobacillus buchneri and Clostridium perfringens, respectively. The L. buchneri enzyme also contained a peptide homologous to the class 1 cysteinyl peptide from Lactobacillus 30a. However, no corresponding peptide was present in the enzyme from C. perfringens, in which the second cysteinyl residue of the alpha chain occupies position 3, very near the essential pyruvoyl residue. This enzyme, unlike those from Lactobacillus 30a or L. buchneri, also contains one cysteinyl residue in its beta chain. Although Cys 147 is an active site residue in histidine decarboxylase from Lactobacillus 30a, the absence of a corresponding residue in the C. perfringens enzyme confirms previous indications (Recsei, P. A., and Snell, E. E. (1982) J. Biol. Chem. 257, 7196-7202) that this SH group is not essential for decarboxylase action.  相似文献   

17.
The biotin-containing oxaloacetate decarboxylase from Klebsiella aerogenes catalyzed the Na+-dependent decarboxylation of oxaloacetate to pyruvate and bicarbonate (or CO2) but not the reversal of this reaction, not even in the presence of an oxaloacetate trapping system. The enzyme catalyzed an avidin-sensitive isotopic exchange between [1-14C]pyruvate and oxaloacetate, which indicated the intermediate formation of a carboxybiotin enzyme. Sodium ions were not required for this partial reaction, but promoted the second partial reaction, the decarboxylation of the carboxybiotin enzyme, thus accounting for the Na+ requirement of the overall reaction. Therefore, the 14CO2-enzyme which was formed upon incubation of the decarboxylase with [4-15C]oxaloacetate, could only be isolated if Na+ ions were excluded. Preincubation of the decarboxylase with avidin also prevented its labelling with 14CO2. The isolated 14CO2-labelled oxaloacetate decarboxylase revealed the following properties. It was slowly decarboxylated at neutral pH and rapidly upon acidification. The 14CO2 residues of the 14CO2-enzyme could be transferred to pyruvate yielding [4-14C]oxaloacetate. In the presence of Na+ this 14CO2 transfer was repressed by the simultaneous decarboxylation of the 14CO2-enzyme. However, Na+ alone was insufficient as a cofactor for the decarboxylation of the isolated 14CO2-enzyme, since this required pyruvate in addition to Na+. It is therefore concluded that the decarboxylation of oxaloacetate proceeds over a CO2-enzyme--pyruvate complex and that free CO2-enzyme is an abortive reaction intermediate. The activation energy of the enzymic decarboxylation of oxaloacetate changed with temperature and was about 113 kJ below 11 degrees C, 60 kJ between 11 degrees C and 31 degrees C and 36 kJ between 31--45 degrees C.  相似文献   

18.
The crenarchaeon Sulfolobus solfataricus uses arginine to produce putrescine for polyamine biosynthesis. However, genome sequences from S. solfataricus and most crenarchaea have no known homologs of the previously characterized pyridoxal 5'-phosphate or pyruvoyl-dependent arginine decarboxylases that catalyze the first step in this pathway. Instead they have two paralogs of the S-adenosylmethionine decarboxylase (AdoMetDC). The gene at locus SSO0585 produces an AdoMetDC enzyme, whereas the gene at locus SSO0536 produces a novel arginine decarboxylase (ArgDC). Both thermostable enzymes self-cleave at conserved serine residues to form amino-terminal beta-domains and carboxyl-terminal alpha-domains with reactive pyruvoyl cofactors. The ArgDC enzyme specifically catalyzed arginine decarboxylation more efficiently than previously studied pyruvoyl enzymes. alpha-Difluoromethylarginine significantly reduced the ArgDC activity of purified enzyme, and treating growing S. solfataricus cells with this inhibitor reduced the cells' ratio of spermidine to norspermine by decreasing the putrescine pool. The crenarchaeal ArgDC had no AdoMetDC activity, whereas its AdoMetDC paralog had no ArgDC activity. A chimeric protein containing the beta-subunit of SSO0536 and the alpha-subunit of SSO0585 had ArgDC activity, implicating residues responsible for substrate specificity in the amino-terminal domain. This crenarchaeal ArgDC is the first example of alternative substrate specificity in the AdoMetDC family. ArgDC activity has evolved through convergent evolution at least five times, demonstrating the utility of this enzyme and the plasticity of amino acid decarboxylases.  相似文献   

19.
Phosphopantothenoylcysteine decarboxylase catalyzes the decarboxylation of 4'-phosphopantothenoylcysteine (2) to form 4'-phosphopanthetheine (3), an intermediate in the biosynthesis of Coenzyme A. In this study we investigated the stereochemistry of this reaction. Our results show that the decarboxylation proceeds with retention of stereochemistry, and that the pro-R proton at C(beta) of the cysteine moiety of 2 is removed during a reversible oxidation of the thiol to a thioaldehyde intermediate.  相似文献   

20.
Photosystem II (PSII), the thylakoid membrane enzyme which uses sunlight to oxidize water to molecular oxygen, holds many organic and inorganic redox cofactors participating in the electron transfer reactions. Among them, two tyrosine residues, Tyr-Z and Tyr-D are found on the oxidizing side of PSII. Both tyrosines demonstrate similar spectroscopic features while their kinetic characteristics are quite different. Tyr-Z, which is bound to the D1 core protein, acts as an intermediate in electron transfer between the primary donor, P(680) and the CaMn? cluster. In contrast, Tyr-D, which is bound to the D2 core protein, does not participate in linear electron transfer in PSII and stays fully oxidized during PSII function. The phenolic oxygens on both tyrosines form well-defined hydrogen bonds to nearby histidine residues, His(Z) and His(D) respectively. These hydrogen bonds allow swift and almost activation less movement of the proton between respective tyrosine and histidine. This proton movement is critical and the phenolic proton from the tyrosine is thought to toggle between the tyrosine and the histidine in the hydrogen bond. It is found towards the tyrosine when this is reduced and towards the histidine when the tyrosine is oxidized. The proton movement occurs at both room temperature and ultra low temperature and is sensitive to the pH. Essentially it has been found that when the pH is below the pK(a) for respective histidine the function of the tyrosine is slowed down or, at ultra low temperature, halted. This has important consequences for the function also of the CaMn? complex and the protonation reactions as the critical Tyr-His hydrogen bond also steer a multitude of reactions at the CaMn? cluster. This review deals with the discovery and functional assignments of the two tyrosines. The pH dependent phenomena involved in oxidation and reduction of respective tyrosine is covered in detail. This article is part of a Special Issue entitled: Photosystem II.  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