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.

Background  

Arabidopsis thaliana transthyretin-like (TTL) protein is a potential substrate in the brassinosteroid signalling cascade, having a role that moderates plant growth. Moreover, sequence homology revealed two sequence domains similar to 2-oxo-4-hydroxy-4-carboxy-5-ureidoimidazoline (OHCU) decarboxylase (N-terminal domain) and 5-hydroxyisourate (5-HIU) hydrolase (C-terminal domain). TTL is a member of the transthyretin-related protein family (TRP), which comprises a number of proteins with sequence homology to transthyretin (TTR) and the characteristic C-terminal sequence motif Tyr-Arg-Gly-Ser. TRPs are single domain proteins that form tetrameric structures with 5-HIU hydrolase activity. Experimental evidence is fundamental for knowing if TTL is a tetrameric protein, formed by the association of the 5-HIU hydrolase domains and, in this case, if the structural arrangement allows for OHCU decarboxylase activity. This work reports about the biochemical and functional characterization of TTL.  相似文献   

2.
Li T  Iwaki H  Fu R  Hasegawa Y  Zhang H  Liu A 《Biochemistry》2006,45(21):6628-6634
The enzymatic activity of Pseudomonas fluorescens alpha-amino-beta-carboxymuconic-epsilon-semialdehyde decarboxylase (ACMSD) is critically dependent on a transition metal ion [Li, T., Walker, A. L., Iwaki, H., Hasegawa, Y., and Liu, A. (2005) J. Am. Chem. Soc. 127, 12282-12290]. Sequence analysis in this study further suggests that ACMSD belongs to the amidohydrolase superfamily, whose structurally characterized members comprise a catalytically essential metal cofactor. To identify ACMSD's metal ligands and assess their functions in catalysis, a site-directed mutagenesis analysis was conducted. Alteration of His-9, His-177, and Asp-294 resulted in a dramatic loss of enzyme activity, substantial reduction of the metal-binding ability, and an altered metallocenter electronic structure. Thus, these residues are confirmed to be the endogenous metal ligands. His-11 is implicated in metal binding because of the strictly conserved HxH motif with His-9. Mutations at the 228 site yielded nearly inactive enzyme variants H228A and H228E. The two His-228 mutant proteins, however, exhibited full metal-binding ability and a metal center similar to that of the wild-type enzyme as shown by EPR spectroscopy. Kinetic analysis on the mutants indicates that His-228 is a critical catalytic residue along with the metal cofactor. Since the identified metal ligands and His-228 are present in all known ACMSD sequences, it is likely that ACMSD proteins from other organisms contain the same cofactor and share similar catalytic mechanisms. ACMSD is therefore the first characterized member in the amidohydrolase superfamily that represents a C-C breaking activity.  相似文献   

3.
Mevalonate (MVA) metabolism provides the isoprenoids used in archaeal lipid biosynthesis. In synthesis of isopentenyl diphosphate, the classical MVA pathway involves decarboxylation of mevalonate diphosphate, while an alternate pathway has been proposed to involve decarboxylation of mevalonate monophosphate. To identify the enzymes responsible for metabolism of mevalonate 5-phosphate to isopentenyl diphosphate in Haloferax volcanii, two open reading frames (HVO_2762 and HVO_1412) were selected for expression and characterization. Characterization of these proteins indicated that one enzyme is an isopentenyl phosphate kinase that forms isopentenyl diphosphate (in a reaction analogous to that of Methanococcus jannaschii MJ0044). The second enzyme exhibits a decarboxylase activity that has never been directly attributed to this protein or any homologous protein. It catalyzes the synthesis of isopentenyl phosphate from mevalonate monophosphate, a reaction that has been proposed but never demonstrated by direct experimental proof, which is provided in this account. This enzyme, phosphomevalonate decarboxylase (PMD), exhibits strong inhibition by 6-fluoromevalonate monophosphate but negligible inhibition by 6-fluoromevalonate diphosphate (a potent inhibitor of the classical mevalonate pathway), reinforcing its selectivity for monophosphorylated ligands. Inhibition by the fluorinated analog also suggests that the PMD utilizes a reaction mechanism similar to that demonstrated for the classical MVA pathway decarboxylase. These observations represent the first experimental demonstration in H. volcanii of both the phosphomevalonate decarboxylase and isopentenyl phosphate kinase reactions that are required for an alternate mevalonate pathway in an archaeon. These results also represent, to our knowledge, the first identification and characterization of any phosphomevalonate decarboxylase.  相似文献   

4.
Thiamin diphosphate-dependent decarboxylases catalyze the non-oxidative decarboxylation of 2-keto carboxylic acids. Although they display relatively low sequence similarity, and broadly different range of substrates, these enzymes show a common homotetrameric structure. Here we describe a kinetic characterization of the substrate spectrum of a recently identified member of this class, the branched chain 2-keto acid decarboxylase (KdcA) from Lactococcus lactis. In order to understand the structural basis for KdcA substrate recognition we developed a homology model of its structure. Ser286, Phe381, Val461 and Met358 were identified as residues that appeared to shape the substrate binding pocket. Subsequently, site-directed mutagenesis was carried out on these residues with a view to converting KdcA into a pyruvate decarboxylase. The results show that the mutations all lowered the Km value for pyruvate and both the S286Y and F381W variants also had greatly increased values of k(cat) with pyruvate as a substrate.  相似文献   

5.
The crystal structure of the flavoprotein Pad1 from Escherichia coli O157:H7 complexed with the cofactor FMN has been determined by the multiple anomalous diffraction method and refined at 2.0 A resolution. This protein is a paralog of UbiX (3-octaprenyl-4-hydroxybenzoate carboxylyase, 51% sequence identity) that catalyzes the third step in ubiquinone biosynthesis and to Saccharomyces cerevisiae Pad1 (54% identity), an enzyme that confers resistance to the antimicrobial compounds phenylacrylic acids through decarboxylation of these compounds. Each Pad1 monomer consists of a typical Rossmann fold containing a non-covalently bound molecule of FMN. The fold of Pad1 is similar to MrsD, an enzyme associated with lantibiotic synthesis; EpiD, a peptidyl-cysteine decarboxylase; and AtHAL3a, the enzyme, which decarboxylates 4'-phosphopantothenoylcysteine to 4'-phosphopantetheine during coenzyme A biosynthesis, all with a similar location of the FMN binding site at the interface between two monomers, yet each having little sequence similarity to one another. All of these proteins associate into oligomers, with a trimer forming the common structural unit in each case. In MrsD and EpiD, which belong to the homo-dodecameric flavin-containing cysteine decarboxylase (HFCD) family, these trimers associate further into dodecamers. Pad1 also forms dodecamers, although the association of the trimers is completely different, resulting in exposure of a different side of the trimer unit to the solvent. This exposure affects the location of the substrate binding site and, specifically, its access to the FMN cofactor. Therefore, Pad1 forms a separate family, distinguishable from the HFCD family.  相似文献   

6.
Huo L  Fielding AJ  Chen Y  Li T  Iwaki H  Hosler JP  Chen L  Hasegawa Y  Que L  Liu A 《Biochemistry》2012,51(29):5811-5821
The previously reported crystal structures of α-amino-β-carboxymuconate-ε-semialdehyde decarboxylase (ACMSD) show a five-coordinate Zn(II)(His)(3)(Asp)(OH(2)) active site. The water ligand is H-bonded to a conserved His228 residue adjacent to the metal center in ACMSD from Pseudomonas fluorescens (PfACMSD). Site-directed mutagenesis of His228 to tyrosine and glycine in this study results in a complete or significant loss of activity. Metal analysis shows that H228Y and H228G contain iron rather than zinc, indicating that this residue plays a role in the metal selectivity of the protein. As-isolated H228Y displays a blue color, which is not seen in wild-type ACMSD. Quinone staining and resonance Raman analyses indicate that the blue color originates from Fe(III)-tyrosinate ligand-to-metal charge transfer. Co(II)-substituted H228Y ACMSD is brown in color and exhibits an electron paramagnetic resonance spectrum showing a high-spin Co(II) center with a well-resolved (59)Co (I = 7/2) eight-line hyperfine splitting pattern. The X-ray crystal structures of as-isolated Fe-H228Y (2.8 ?) and Co-substituted (2.4 ?) and Zn-substituted H228Y (2.0 ? resolution) support the spectroscopic assignment of metal ligation of the Tyr228 residue. The crystal structure of Zn-H228G (2.6 ?) was also determined. These four structures show that the water ligand present in WT Zn-ACMSD is either missing (Fe-H228Y, Co-H228Y, and Zn-H228G) or disrupted (Zn-H228Y) in response to the His228 mutation. Together, these results highlight the importance of His228 for PfACMSD's metal specificity as well as maintaining a water molecule as a ligand of the metal center. His228 is thus proposed to play a role in activating the metal-bound water ligand for subsequent nucleophilic attack on the substrate.  相似文献   

7.
DNA methylation and demethylation regulate many crucial biological processes in mammals and are linked to many diseases. Active DNA demethylation is believed to be catalyzed by TET proteins and a putative DNA decarboxylase that may share some similarities in sequence, structure and catalytic mechanism with isoorotate decarboxylase (IDCase) that catalyzes decarboxylation of 5caU to U in fungi. We report here the structures of wild-type and mutant IDCases from Cordyceps militaris and Metarhizium anisopliae in apo form or in complexes with 5caU, U, and an inhibitor 5-nitro-uracil. IDCases adopt a typical (β/α)8 barrel fold of the amidohydrolase superfamily and function as dimers. A Zn2+ is bound at the active site and coordinated by four strictly conserved residues, one Asp and three His. The substrate is recognized by several strictly conserved residues. The functional roles of the key residues at the active site are validated by mutagenesis and biochemical studies. Based on the structural and biochemical data, we present for the first time a novel catalytic mechanism of decarboxylation for IDCases, which might also apply to other members of the amidohydrolase superfamily. In addition, our biochemical data show that IDCases can catalyze decarboxylation of 5caC to C albeit with weak activity, which is the first in vitro evidence for direct decarboxylation of 5caC to C by an enzyme. These findings are valuable in the identification of potential DNA decarboxylase in mammals.  相似文献   

8.
Orotidine 5'-monophoshate decarboxylase (OMPDC) catalyses the decarboxylation of orotidine 5'-monophosphate (OMP) to uridine 5'-monophosphate (UMP). Here, we report the X-ray analysis of apo, substrate or product-complex forms of OMPDC from Plasmodium falciparum (PfOMPDC) at 2.7, 2.65 and 2.65 A, respectively. The structural analysis provides the substrate recognition mechanism with dynamic structural changes, as well as the rearrangement of the hydrogen bond array at the active site. The structural basis of substrate or product binding to PfOMPDC will help to uncover the decarboxylation mechanism and facilitate structure-based optimization of antimalarial drugs.  相似文献   

9.
10.
Genes for two enzymes in the thymidine salvage pathway, thymine-7-hydroxylase (THase; official name thymine dioxygenase) and iso-orotate decarboxylase (IDCase) have been isolated from fungal sources. THase was isolated from a Rhodotorula glutinis cDNA library using a degenerate oligonucleotide based on the published amino acid sequence. The coding sequence was transferred to an Escherichia coli expression system, from which recombinant THase activity was measured using 14C-labeled thymine. The THase sequence shows an almost complete avoidance of codons ending in A or T: 95.8% GC content is present in the third position of codons. A connection between this codon bias and the role of the thymidine salvage pathway in pyrimidine metabolism is proposed. The THase sequence is similar to Group I Fe+2-dependent, alphaKG-dependent dioxygenases. The R. glutinis THase gene was used to locate the probable THase genes in the sequenced genomes of Neurospora crassa and Aspergillus nidulans. The genes neighboring THase in these two genomes are similar to each other, and are similar to the mammalian 2-amino-3-carboxymuconate-6-semialdhyde decarboxylase (ACMSD), leading to their identification as IDCase genes. The N. crassa version was isolated by PCR of genomic DNA, and IDCase activity was measured in recombinant E. coli carrying this gene. A new family of decarboxylases, using similar substrates, is identified by virtue of the protein sequence similarity.  相似文献   

11.
Quinolinate (quinolinic acid) is a potent endogenous excitotoxin of neuronal cells. Elevation of quinolinate levels in the brain has been implicated in the pathogenesis of various neurodegenerative disorders, the so-called "quinolinate hypothesis." Quinolinate is non-enzymatically derived from alpha-amino-beta-carboxymuconate-epsilon-semialdehyde (ACMS). Alpha-amino-beta-carboxymuconate-epsilon-semialdehyde decarboxylase (ACMSD) is the only known enzyme that can process ACMS to a benign catabolite and thus prevent the accumulation of quinolinate from ACMS. ACMSD seems to be regulated by nutritional and hormonal signals, but its molecular mechanism has, to date, been largely unknown. Utilizing partial amino acid sequences obtained from highly purified porcine kidney ACMSD, a cDNA encoding human ACMSD was cloned and characterized. The cDNA encodes a unique open reading frame of 336 amino acids and displays little homology to any known enzymes or motifs in mammalian databases, suggesting that ACMSD may contain a new kind of protein fold. Real-time PCR-based quantification of ACMSD revealed very low but significant levels of the expression in the brain. Brain ACMSD messages were down- and up-regulated in response to low protein diet and streptozocin-induced diabetes, respectively. The enzyme activities measured from partially purified brains were closely correlated with the changes in the message levels. Expression of quinolinate phosphoribosyltransferase (QPRT), another enzyme that catabolizes quinolinate, was also found in the brain. This suggests that a pathway does exist by which the levels of quinolinate in the brain are regulated. In this report, we address the molecular basis underlying quinolinate metabolism and the regulation of ACMSD expression.  相似文献   

12.
The nonoxidative decarboxylation of aromatic acids occurs in a range of microbes and is of interest for bioprocessing and metabolic engineering. Although phenolic acid decarboxylases provide useful tools for bioindustrial applications, the molecular bases for how these enzymes function are only beginning to be examined. Here we present the 2.35-Å-resolution X-ray crystal structure of the ferulic acid decarboxylase (FDC1; UbiD) from Saccharomyces cerevisiae. FDC1 shares structural similarity with the UbiD family of enzymes that are involved in ubiquinone biosynthesis. The position of 4-vinylphenol, the product of p-coumaric acid decarboxylation, in the structure identifies a large hydrophobic cavity as the active site. Differences in the β2e-α5 loop of chains in the crystal structure suggest that the conformational flexibility of this loop allows access to the active site. The structure also implicates Glu285 as the general base in the nonoxidative decarboxylation reaction catalyzed by FDC1. Biochemical analysis showed a loss of enzymatic activity in the E285A mutant. Modeling of 3-methoxy-4-hydroxy-5-decaprenylbenzoate, a partial structure of the physiological UbiD substrate, in the binding site suggests that an ∼30-Å-long pocket adjacent to the catalytic site may accommodate the isoprenoid tail of the substrate needed for ubiquinone biosynthesis in yeast. The three-dimensional structure of yeast FDC1 provides a template for guiding protein engineering studies aimed at optimizing the efficiency of aromatic acid decarboxylation reactions in bioindustrial applications.  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.
Malonate semialdehyde decarboxylase (MSAD) has been identified as the protein encoded by the orf130 gene from Pseudomonas pavonaceae 170 on the basis of the genomic context of the gene as well as its ability to catalyze the decarboxylation of malonate semialdehyde to generate acetaldehyde. The enzyme is found in a degradative pathway for the xenobiotic nematocide trans-1,3-dichloropropene. MSAD has no sequence homology to previously characterized decarboxylases, but the presence of a conserved motif (Pro1-(X)8 -Gly-Arg11-X-Asp-X-Gln) in its N-terminal region suggested a relationship to the tautomerase superfamily. Sequence analysis identified Pro1 and Arg75 as potential active site residues that might be involved in the MSAD activity. The results of site-directed mutagenesis experiments confirmed the importance of these residues to activity and provided further evidence to implicate MSAD as a new member of the tautomerase superfamily. MSAD is the first identified decarboxylase in the superfamily and is possibly the first characterized member of a new and distinct family within this superfamily. Malonate semialdehyde is analogous to a beta-keto acid, and enzymes that catalyze the decarboxylation of these acids generally utilize metal ion catalysis, a Schiff base intermediate, or polarization of the carbonyl group by hydrogen bonding and/or electrostatic interactions. A mechanistic analysis shows that the rate of the reaction is not affected by the presence of a metal ion or EDTA while the incubation of MSAD with the substrate in the presence of sodium cyanoborohydride results in the irreversible inactivation of the enzyme. The site of modification is Pro1. These observations are consistent with the latter two mechanisms, but do not exclude the first mechanism. Based on the sequence analysis, the outcome of the mutagenesis and mechanistic experiments, and the roles determined for Pro1 and the conserved arginine in all tautomerase superfamily members characterized thus far, two mechanistic scenarios are proposed for the MSAD-catalyzed reaction in which Pro1 and Arg75 play prominent roles.  相似文献   

15.
16.
Phosphorylation and decarboxylation of mevalonate in chick liver and brain was investigated during early post hatching stages of development. In chick liver, both mevalonate kinase and mevalonate-5-phosphate kinase increased their activity from day 5 of age while pyrophosphate decarboxylase activity remained low during the first days after hatching, increased sharply up to day 9 of age, and remained practically unchanged thereafter. The developmental pattern obtained in brain shows a slight decrease in the phosphorylation and decarboxylation of mevalonate after the first week of postnatal development. Further studies were performed using the specific substrate of mevalonate-5-pyrophosphate decarboxylase, corroborating the results obtained using mevalonate as substrate. Changes in hepatic decarboxylase were more pronounced than those observed in mevalonate-phosphorylating enzymes, thus suggesting an important role for decarboxylase in the control of cholesterogenesis during postnatal development.  相似文献   

17.
A major pathway of beta-alanine synthesis in insects is through the alpha-decarboxylation of aspartate, but the enzyme involved in the decarboxylation of aspartate has not been clearly defined in mosquitoes and characterized in any insect species. In this study, we expressed two putative mosquito glutamate decarboxylase-like enzymes of mosquitoes and critically analyzed their substrate specificity and biochemical properties. Our results provide clear biochemical evidence establishing that one of them is an aspartate decarboxylase and the other is a glutamate decarboxylase. The mosquito aspartate decarboxylase functions exclusively on the production of beta-alanine with no activity with glutamate. Likewise the mosquito glutamate decarboxylase is highly specific to glutamate with essentially no activity with aspartate. Although insect aspartate decarboxylase shares high sequence identity with glutamate decarboxylase, we are able to closely predict aspartate decarboxylase from glutamate decarboxylase based on the difference of their active site residues.  相似文献   

18.
Pyruvate decarboxylases (PDCs) are a class of enzymes which carry out the non-oxidative decarboxylation of pyruvate to acetaldehyde. These enzymes are also capable of carboligation reactions and can generate chiral intermediates of substantial pharmaceutical interest. Typically, the decarboxylation and carboligation processes are carried out using whole cell systems. However, fermentative organisms such as Saccharomyces cerevisiae are known to contain several PDC isozymes; the precise suitability and role of each of these isozymes in these processes is not well understood. S. cerevisiae has three catalytic isozymes of pyruvate decarboxylase (ScPDCs). Of these, ScPDC1 has been investigated in detail by various groups with the other two catalytic isozymes, ScPDC5 and ScPDC6 being less well characterized. Pyruvate decarboxylase activity can also be detected in the cell lysates of Komagataella pastoris, a Crabtree-negative yeast, and consequently it is of interest to investigate whether this enzyme has different kinetic properties. This is also the first report of the expression and functional characterization of pyruvate decarboxylase from K. pastoris (PpPDC). This investigation helps in understanding the roles of the three isozymes at different phases of S. cerevisiae fermentation as well as their relevance for ethanol and carboligation reactions. The kinetic and physical properties of the four isozymes were determined using similar conditions of expression and characterization. ScPDC5 has comparable decarboxylation efficiency to that of ScPDC1; however, the former has the highest rate of reaction, and thus can be used for industrial production of ethanol. ScPDC6 has the least decarboxylation efficiency of all three isozymes of S. cerevisiae. PpPDC in comparison to all isozymes of S. cerevisiae is less efficient at decarboxylation. All the enzymes exhibit allostery, indicating that they are substrate activated.  相似文献   

19.
Rishavy MA  Cleland WW 《Biochemistry》2000,39(16):4569-4574
Orotidine 5'-monophosphate shows a (15)N isotope effect of 1.0036 at N-1 for decarboxylation catalyzed by orotidine 5'-monophosphate decarboxylase. Picolinic acid shows a (15)N isotope effect of 0.9955 for decarboxylation in ethylene glycol at 190 degrees C, while N-methyl picolinic acid shows a (15)N isotope effect of 1.0053 at 120 degrees C. The transition state for enzymatic decarboxylation of orotidine 5'-monophosphate resembles the transition state for N-methyl picolinic acid in that no bond order changes take place at N-1. This rules out enolization to give a quaternary nitrogen at N-1 in the enzymatic mechanism and suggests a carbanion intermediate stabilized by simple electrostatic interaction with Lys-93. The driving force for the reaction appears to be ground-state destabilization resulting from charge repulsion between the carboxyl of the substrate and Asp-91.  相似文献   

20.
4-Oxalocrotonate decarboxylase (4-OD) and vinylpyruvate hydratase (VPH) from Pseudomonas putida mt-2 form a complex that converts 2-oxo-3-hexenedioate to 2-oxo-4-hydroxypentanoate in the catechol meta fission pathway. To facilitate mechanistic and structural studies of the complex, the two enzymes have been coexpressed and the complex has been purified to homogeneity. In addition, Glu-106, a potential catalytic residue in VPH, has been changed to glutamine, and the resulting E106QVPH mutant has been coexpressed with 4-OD and purified to homogeneity. The 4-OD/E106QVPH complex retains full decarboxylase activity, with comparable kinetic parameters to those observed for 4-OD in the wild-type complex, but is devoid of any detectable hydratase activity. Decarboxylation of (5S)-2-oxo-3-[5-D]hexenedioate by either the 4-OD/VPH complex or the mutant complex generates 2-hydroxy-2,4E-[5-D]pentadienoate in D(2)O. Ketonization of 2-hydroxy-2,4-pentadienoate by the wild-type complex is highly stereoselective and results in the formation of 2-oxo-(3S)-[3-D]-4-pentenoate, while the mutant complex generates a racemic mixture. These results indicate that 2-hydroxy-2, 4-pentadienoate is the product of 4-OD and that 2-oxo-4-pentenoate results from a VPH-catalyzed process. On this basis, the previously proposed hypothesis for the conversion of 2-oxo-3-hexenedioate to 2-oxo-4-hydroxypentanoate has been revised [Lian, H., and Whitman, C. P. (1994) J. Am. Chem. Soc. 116, 10403-10411]. Finally, the observed (13)C kinetic isotope effect on the decarboxylation of 2-oxo-3-hexenedioate by the 4-OD/VPH complex suggests that the decarboxylation step is nearly rate-limiting. Because the value is not sensitive to either magnesium or manganese, it is likely that the transition state for carbon-carbon bond cleavage is late and that the metal positions the substrate and polarizes the carbonyl group, analogous to its role in oxalacetate decarboxylase.  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