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.
A homology model for the pig isozyme of the pyridoxal phosphate-dependent enzyme gamma-aminobutyrate (GABA) aminotransferase has been built based mainly on the structure of dialkylglycine decarboxylase and on a multiple sequence alignment of 28 evolutionarily related enzymes. The proposed active site structure is presented and analyzed. Hypothetical structures for external aldimine intermediates explain several characteristics of the enzyme. In the GABA external aldimine model, the pro-S proton at C4 of GABA, which abstracted in the 1,3-azaallylic rearrangement interconverting the aldimine and ketimine intermediates, is oriented perpendicular to the plane of the pyridoxal phosphate ring. Lys 329 is in close proximity and is probably the general base catalyst for the proton transfer reaction. The carboxylate group of GABA interacts with Arg 192 and Lys 203, which determine the specificity of the enzyme for monocarboxylic omega-amino acids such as GABA. In the proposed structure for the L-glutamate external aldimine, the alpha-carboxylate interacts with Arg 445. Glu 265 is proposed to interact with this same arginine in the GABA external aldimine, enabling the enzyme to act on omega-amino acids in one half-reaction and on alpha-amino acids in the other. The reactivities of inhibitors are well explained by the proposed active site structure. The R and S isomers of beta-substituted phenyl and p-chlorophenyl GABA would bind in very different modes due to differential steric interactions, with the reactive S isomer leaving the orientation of the GABA moiety relatively unperturbed compared to that of the natural substrate. In our model, only the reactive S isomer of the mechanism-based inhibitor vinyl-GABA, an effective anti-epileptic drug known clinically as Vigabatrin, would orient the scissile C4-H bond perpendicular to the coenzyme ring plane and present the proton to Lys 329, the proposed general base catalyst of the reaction. The R isomer would direct the vinyl group toward Lys 329 and the C4-H bond toward Arg 445. The active site model presented provides a basis for site-directed mutagenesis and drug design experiments.  相似文献   

2.
Biosynthetic ornithine decarboxylase was purified 4300-fold from Escherichia coli to a purity of approximately 85% as judged by polyacrylamide gel electrophoresis. The enzyme showed hyperbolic kinetics with a Km of 5.6 mM for ornithine and 1.0 micronM for pyridoxal phosphate and it was competitively inhibited by putrescine and spermidine. The biosynthetic decarboxylase was compared with the biodegradative ornithine decarboxylase [Applebaum, D., et al. (1975), Biochemistry 14, 3675]. Both enzymes were dimers of 80 000-82 000 molecular weight and exhibited similar kinetic properties. However, they differed significantly in other respects. The pH optimum of the biosynthetic enzyme was 8.1, compared with 6.9 for the biodegradative. Both enzymes were activated by nucleotides, but with different specificity. Antibody to the purified biodegradative ornithine decarboxylase did not cross-react with the biosynthetic enzyme. The evolutionary relationship of these two decarboxylases to the other amino acid decarboxylases of E. coli is discussed.  相似文献   

3.
1-Aminooxy-3-aminopropane was shown to be a potent competitive inhibitor (Ki = 3.2 nM) of homogenous mouse kidney ornithine decarboxylase, a potent irreversible inhibitor (Ki = 50 microM) of homogeneous liver adenosylmethionine decarboxylase and a potent competitive (Ki = 2.3 microM) of homogeneous bovine brain spermidine synthase. It did not inhibit homogeneous bovine brain spermine synthase and it did not serve as a substrate for spermidine synthase. The compound did not inhibit tyrosine aminotransferase, alanine aminotransferase or aspartate aminotransferase, which are pyridoxal phosphate-containing enzymes like ornithine decarboxylase. The inactivation of adenosylmethionine decarboxylase was partially prevented by pyruvate, which is the coenzyme of adenosylmethionine decarboxylase, and by the substrate, adenosylmethionine. 1-Aminooxy-3-aminopropane at 0.5 mM concentration inhibited the growth of HL-60 promyelocytic leukemia cells and this inhibition was prevented by spermidine but not by putrescine.  相似文献   

4.
The pyridoxal-P binding sites of the two isoforms of human glutamate decarboxylase (GAD65 and GAD67) were modeled by using PROBE (a recently developed algorithm for multiple sequence alignment and database searching) to align the primary sequence of GAD with pyridoxal-P binding proteins of known structure. GAD's cofactor binding site is particularly interesting because GAD activity in the brain is controlled in part by a regulated interconversion of the apo- and holoenzymes. PROBE identified six motifs shared by the two GADs and four proteins of known structure: bacterial ornithine decarboxylase, dialkylglycine decarboxylase, aspartate aminotransferase, and tyrosine phenol-lyase. Five of the motifs corresponded to the alpha/beta elements and loops that form most of the conserved fold of the pyridoxal-P binding cleft of the four enzymes of known structure; the sixth motif corresponded to a helical element of the small domain that closes when the substrate binds. Eight residues that interact with pyridoxal-P and a ninth residue that lies at the interface of the large and small domains were also identified. Eleven additional conserved residues were identified and their functions were evaluated by examining the proteins of known structure. The key residues that interact directly with pyridoxal-P were identical in ornithine decarboxylase and the two GADs, thus allowing us to make a specific structural prediction of the cofactor binding site of GAD. The strong conservation of the cofactor binding site in GAD indicates that the highly regulated transition between apo- and holoGAD is accomplished by modifications in this basic fold rather than through a novel folding pattern.  相似文献   

5.
Polyamines are fundamental molecules of life, and their deep evolutionary history is reflected in extensive biosynthetic diversification. The polyamines putrescine, agmatine, and cadaverine are produced by pyridoxal 5′-phosphate-dependent L-ornithine, L-arginine, and L-lysine decarboxylases (ODC, ADC, LDC), respectively, from both the alanine racemase (AR) and aspartate aminotransferase (AAT) folds. Two homologous forms of AAT-fold decarboxylase are present in bacteria: an ancestral form and a derived, acid-inducible extended form containing an N-terminal fusion to the receiver-like domain of a bacterial response regulator. Only ADC was known from the ancestral form and limited to the Firmicutes phylum, whereas extended forms of ADC, ODC, and LDC are present in Proteobacteria and Firmicutes. Here, we report the discovery of ancestral form ODC, LDC, and bifunctional O/LDC and extend the phylogenetic diversity of functionally characterized ancestral ADC, ODC, and LDC to include phyla Fusobacteria, Caldiserica, Nitrospirae, and Euryarchaeota. Using purified recombinant enzymes, we show that these ancestral forms have a nascent ability to decarboxylate kinetically less preferred amino acid substrates with low efficiency, and that product inhibition primarily affects preferred substrates. We also note a correlation between the presence of ancestral ODC and ornithine/arginine auxotrophy and link this with a known symbiotic dependence on exogenous ornithine produced by species using the arginine deiminase system. Finally, we show that ADC, ODC, and LDC activities emerged independently, in parallel, in the homologous AAT-fold ancestral and extended forms. The emergence of the same ODC, ADC, and LDC activities in the nonhomologous AR-fold suggests that polyamine biosynthesis may be inevitable.  相似文献   

6.
Ornithine decarboxylase has been purified 1,500-fold to homogeneity from a spe2 mutant of Saccharomyces cerevisiae which lacks S-adenosylmethionine decarboxylase and is derepressed for ornithine decarboxylase. The ornithine decarboxylase is a single polypeptide (Mr = 68,000) and requires a thiol and pyridoxal phosphate for activity. Addition of 10(-4) M spermidine and 10(-4) M spermine to the growth medium reduces the activity of the enzyme by 90% in 4 h. However, immunoprecipitation studies showed that the extracts of polyamine-treated cells contain as much enzyme protein as normal cell extracts. This loss of ornithine decarboxylase activity is probably due to a post-translational modification of enzyme protein because we found no evidence for any inhibitor of activity in the polyamine-treated cells.  相似文献   

7.
The genome sequence of the hyperthermophilic methanogen Methanococcus jannaschii contains homologs of most genes required for spermidine polyamine biosynthesis. Yet genomes from neither this organism nor any other euryarchaeon have orthologs of the pyridoxal 5'-phosphate-dependent ornithine or arginine decarboxylase genes, required to produce putrescine. Instead, as shown here, these organisms have a new class of arginine decarboxylase (PvlArgDC) formed by the self-cleavage of a proenzyme into a 5-kDa subunit and a 12-kDa subunit that contains a reactive pyruvoyl group. Although this extremely thermostable enzyme has no significant sequence similarity to previously characterized proteins, conserved active site residues are similar to those of the pyruvoyl-dependent histidine decarboxylase enzyme, and its subunits form a similar (alphabeta)(3) complex. Homologs of PvlArgDC are found in several bacterial genomes, including those of Chlamydia spp., which have no agmatine ureohydrolase enzyme to convert agmatine (decarboxylated arginine) into putrescine. In these intracellular pathogens, PvlArgDC may function analogously to pyruvoyl-dependent histidine decarboxylase; the cells are proposed to import arginine and export agmatine, increasing the pH and affecting the host cell's metabolism. Phylogenetic analysis of Pvl- ArgDC proteins suggests that this gene has been recruited from the euryarchaeal polyamine biosynthetic pathway to function as a degradative enzyme in bacteria.  相似文献   

8.
Summary A database search has revealed significant and extensive sequence similarities among prokaryotic and eukaryotic pyridoxal phosphate (PLP)-dependent decarboxylases, includingDrosophila glutamic acid decarboxylase (GAD) and bacterial histidine decarboxylase (HDC). Based on these findings, the sequences of seven PLP-dependent decarboxylases from five different organisms have been aligned to derive a consensus sequence for this family of enzymes. In addition, quantitative methods have been employed to calculate the relative evolutionary distances between pairs of the decarboxylases comprising this family. The multiple sequence analysis together with the quantitative results strongly suggest an ancient and common origin for all PLP-dependent decarboxylases. This analysis also indicates that prokaryotic and eukaryotic HDC activities evolved independently. Finally, a sensitive search algorithm (PROFILE) was unable to detect additional members of this decarboxylase family in protein sequence databases.  相似文献   

9.
Comparisons were made of ornithine decarboxylase isolated from Morris hepatoma 7777, thioacetamide-treated rat liver and androgen-stimulated mouse kidney. The enzymes from each source were purified in parallel and their size, isoelectric point, interaction with a monoclonal antibody or a monospecific rabbit antiserum to ornithine decarboxylase, and rates of inactivation in vitro, were studied. Mouse kidney, which is a particularly rich source of ornithine decarboxylase after androgen induction, contained two distinct forms of the enzyme which differed slightly in isoelectric point, but not in Mr. Both forms had a rapid rate of turnover, and virtually all immunoreactive ornithine decarboxylase protein was lost within 4h after protein synthesis was inhibited. Only one form of ornithine decarboxylase was found in thioacetamide-treated rat liver and Morris hepatoma 7777. No differences between the rat liver and hepatoma ornithine decarboxylase protein were found, but the rat ornithine decarboxylase could be separated from the mouse kidney ornithine decarboxylase by two-dimensional gel electrophoresis. The rat protein was slightly smaller and had a slightly more acid isoelectric point. Studies of the inactivation of ornithine decarboxylase in vitro in a microsomal system [Zuretti & Gravela (1983) Biochim. Biophys. Acta 742, 269-277] showed that the enzymes from rat liver and hepatoma 7777 and mouse kidney were inactivated at the same rate. This inactivation was not due to degradation of the enzyme protein, but was probably related to the formation of inactive forms owing to the absence of thiol-reducing agents. Treatment with 1,3-diaminopropane, which is known to cause an increase in the rate of degradation of ornithine decarboxylase in vivo [Seely & Pegg (1983) Biochem. J. 216, 701-717] did not stimulate inactivation by microsomal extracts, indicating that this system does not correspond to the rate-limiting step of enzyme breakdown in vivo.  相似文献   

10.
Myxamoebae of Dictyostelium discoideum from exponentially growing cultures showed altered ornithine decarboxylase activity upon external osmotic perturbation. On transfer to hypotonic NaCl solutions (20 mosmol/kg), cells showed high enzyme activity which was relatively independent of the concentration of the coenzyme pyridoxal phosphate (assay concentrations, 5 and 200 microM). In hypertonic solution (400 mosmol/kg) cells had a reduced level of ornithine decarboxylase activity which was dependent on the coenzyme concentration. The changes in activity were freely reversible in further external osmotic manipulation. The response to osmotic change occurred rapidly, within a few minutes. The changes still occurred at 7 degrees C but 2 mM sodium azide prevented the formation of the high activity form, although this effect was reversed when azide was removed. Cycloheximide had no effect on the osmotically induced changes. Addition of putrescine caused ornithine decarboxylase eventually to the converted to the low-activity form regardless of the osmolality of the solution. The characteristic cofactor concentration dependence of the high- and low-activity form were retained on storage of the cell extracts. No evidence was found for diffusible effectors which stabilized one or the other form of the activity. The enzymes responsible for the two forms were of the same molecular size as judged by gel filtration, and the activities had similar thermostabilities. The results are interpreted in terms of an osmotically induced interconversion of two forms of a single ornithine decarboxylase.  相似文献   

11.
Vitamin B(6)-dependent enzymes may be grouped into five evolutionarily unrelated families, each having a different fold. Within fold type I enzymes, L-threonine aldolase (L-TA) and fungal alanine racemase (AlaRac) belong to a subgroup of structurally and mechanistically closely related proteins, which specialised during evolution to perform different functions. In a previous study, a comparison of the catalytic properties and active site structures of these enzymes suggested that they have a catalytic apparatus with the same basic features. Recently, recombinant D-threonine aldolases (D-TAs) from two bacterial organisms have been characterised, their predicted amino acid sequences showing no significant similarities to any of the known B(6) enzymes. In the present work, a comparative structural analysis suggests that D-TA has an alpha/beta barrel fold and therefore is a fold type III B(6) enzyme, as eukaryotic ornithine decarboxylase (ODC) and bacterial AlaRac. The presence of both TA and AlaRac in two distinct evolutionary unrelated families represents a novel and interesting example of convergent evolution. The independent emergence of the same catalytic properties in families characterised by completely different folds may have not been determined by chance, but by the similar structural features required to catalyse pyridoxal phosphate-dependent aldolase and racemase reactions.  相似文献   

12.
This review considers the role of antizyme, of amino acids and of protein synthesis in the regulation of polyamine biosynthesis.The ornithine decarboxylase of eukaryotic ceils and ofEscherichia coli coli can be non-competitively inhibited by proteins, termed antizymes, which are induced by di-and poly- amines. Some antizymes have been purified to homogeneity and have been shown to be structurally unique to the cell of origin. Yet, the E. c o l i antizyme and the rat liver antizyme cross react and inhibit each other's biosynthetic decarboxylases. These results indicate that aspects of the control of polyamine biosynthesis have been highly conserved throughout evolution.Evidence for the physiological role of the antizyme in mammalian cells rests upon its identification in normal uninduced cells, upon the inverse relationship that exists between antizyme and ornithine decarboxylase as well as upon the existence of the complex of ornithine decarboxylase and antizyme in vivo. Furthermore, the antizyme has been shown to be highly specific; its Keq for ornithine decarboxylase is 1.4 x 1011 M-1. In addition, mammalian ceils contain an anti-antizyme, a protein that specifically binds to the antizyme of an ornithine decarboxylase-antizyme complex and liberates free ornithine decarboxylase from the complex. In B. coli , in which polyamine biosynthesis is mediated both by ornithine decarboxylase and by arginine decarboxylase, three proteins (one acidic and two basic) have been purified, each of which inhibits both these enzymes. They do not inhibit the biodegradative ornithine and arginine decarboxylases nor lysine decarboxylase. The two basic inhibitors have been shown to correspond to the ribosomal proteins S20/L26 and L34, respectively. The relationship of the acidic antizyme to other known B. coli proteins remains to be determined.  相似文献   

13.
1. Polyamine concentrations were decreased in rats fed on a diet deficient in vitamin B-6. 2. Ornithine decarboxylase activity was decreased by vitamin B-6 deficiency when assayed in tissue extracts without addition of pyridoxal phosphate, but was greater than in control extracts when pyridoxal phosphate was present in saturating amounts. 3. In contrast, the activity of S-adenosylmethionine decarboxylase was not enhanced by pyridoxal phosphate addition even when dialysed extracts were prepared from tissues of young rats suckled by mothers fed on the vitamin B-6-deficient diet. 4. S-Adenosylmethionine decarboxylase activities were increased by administration of methylglyoxal bis(guanylhydrazone) (1,1'-[(methylethanediylidine)dinitrilo]diguanidine) to similar extents in both control and vitamin B-6-deficient animals. 5. The spectrum of highly purified liver S-adenosylmethionine decarboxylase did not indicate the presence of pyridoxal phosphate. After inactivation of the enzyme by reaction with NaB3H4, radioactivity was incorporated into the enzyme, but was not present as a reduced derivative of pyridoxal phosphate. 6. It is concluded that the decreased concentrations of polyamines in rats fed on a diet containing vitamin B-6 may be due to decreased activity or ornithine decarboxylase or may be caused by an unknown mechanism responding to growth retardation produced by the vitamin deficiency. In either case, measurements of S-adenosylmethionine decarboxylase and ornithine decarboxylase activity under optimum conditions in vitro do not correlate with the polyamine concentrations in vivo.  相似文献   

14.
15.
Diaminopimelate decarboxylase (DAPDC) and ornithine decarboxylase (ODC) are pyridoxal 5'-phosphate dependent enzymes that are critical to microbial growth and pathogenicity. The latter is the target of drugs that cure African sleeping sickness, while the former is an attractive target for antibacterials. These two enzymes share the (β/α)(8) (i.e., TIM barrel) fold with alanine racemase, another pyridoxal 5'-phosphate dependent enzyme critical to bacterial survival. The active site structural homology between DAPDC and ODC is striking even though DAPDC catalyzes the decarboxylation of a D stereocenter with inversion of configuration and ODC catalyzes the decarboxylation of an L stereocenter with retention of configuration. Here, the structural and mechanistic bases of these interesting properties are explored using reactions of alternate substrates with both enzymes. It is concluded that simple binding determinants do not control the observed stereochemical specificities for decarboxylation, and a concerted decarboxylation/proton transfer at Cα of the D stereocenter of diaminopimelate is a possible mechanism for the observed specificity with DAPDC.  相似文献   

16.
The metabolism of polyamines as well as their functions as growth regulators in plants have been extensively studied for many years. However, almost nothing is known about the biosynthesis and roles of these substances in Phytomonas spp., parasites of several plants. We have used HPLC and electrophoretic analyses to investigate the presence and metabolism of polyamines in Phytomonas Jma strain, detecting both putrescine and spermidine but not spermine. Experiments carried out by incubation of intact parasites with labelled ornithine or putrescine showed the formation of radioactive putrescine or spermidine, respectively. These results indicated that Phytomonas Jma can synthesise these polyamines through the action of ornithine decarboxylase (ODC) and spermidine synthase. On the other hand, we could not detect the conversion of arginine to agmatine, suggesting the absence of arginine decarboxylase (ADC) in Phytomonas. However, we cannot ensure the complete absence of this enzymatic activity in the parasite. Phytomonas ODC required pyridoxal 5′-phosphate for maximum activity and was specifically inhibited by α-difluoromethylornithine. The metabolic turnover of the enzyme was very high, with a half-life of 10-15 min, one of the shortest found among all ODC enzymes studied to date. The parasite proteasome seems to be involved in degradation of the enzyme, since Phytomonas ODC can be markedly stabilized by MG-132, a well known proteasome inhibitor. The addition of polyamines to Phytomonas cultures did not decrease ODC activity, strongly suggesting the possible absence of antizyme in this parasite.  相似文献   

17.
A 3969-base pair PstI-PstI fragment of Pseudomonas cepacia DNA containing the gene for the pyridoxal 5'-phosphate dependent 2,2-dialkylglycine decarboxylase (pyruvate) (EC 4.1.1.64) was cloned in Escherichia coli. The insert was sequenced by the dideoxy method using nested deletions from both ends, revealing a central 1302-base pair region that codes for the decarboxylase subunit. The recombinant enzyme was expressed in E. coli, purified to homogeneity, and sequenced at the amino terminus. Also, a cofactor-labeled active site peptide was sequenced. The carboxyl terminus of the deduced amino acid sequence is homologous with the carboxyl terminus of mammalian ornithine aminotransferase; the active site sequence is similar to the active site sequences of several other aminotransferases. No homologies with known decarboxylase sequences could be found. Expression of the decarboxylase gene is negatively controlled by a 687-nucleotide sequence upstream of and diverging from the structural gene. Expression is induced by S-isovaline, 2-methylalanine, and D-2-aminobutanoic acid, but not by glycine, D- or L-alanine, L-2-aminobutanoic acid, R-isovaline, or other alkyl amino acids.  相似文献   

18.
During the course of infection, Salmonella enterica serovar Typhimurium must successively survive the harsh acid stress of the stomach and multiply into a mild acidic compartment within macrophages. Inducible amino acid decarboxylases are known to promote adaptation to acidic environments. Three low pH inducible amino acid decarboxylases were annotated in the genome of S. Typhimurium, AdiA, CadA and SpeF, which are specific for arginine, lysine and ornithine, respectively. In this study, we characterized and compared the contributions of those enzymes in response to acidic challenges. Individual mutants as well as a strain deleted for the three genes were tested for their ability (i) to survive an extreme acid shock, (ii) to grow at mild acidic pH and (iii) to infect the mouse animal model. We showed that the lysine decarboxylase CadA had the broadest range of activity since it both had the capacity to promote survival at pH 2.3 and growth at pH 4.5. The arginine decarboxylase AdiA was the most performant in protecting S. Typhimurium from a shock at pH 2.3 and the ornithine decarboxylase SpeF conferred the best growth advantage under anaerobiosis conditions at pH 4.5. We developed a GFP-based gene reporter to monitor the pH of the environment as perceived by S. Typhimurium. Results showed that activities of the lysine and ornithine decarboxylases at mild acidic pH did modify the local surrounding of S. Typhimurium both in culture medium and in macrophages. Finally, we tested the contribution of decarboxylases to virulence and found that these enzymes were dispensable for S. Typhimurium virulence during systemic infection. In the light of this result, we examined the genomes of Salmonella spp. normally responsible of systemic infection and observed that the genes encoding these enzymes were not well conserved, supporting the idea that these enzymes may be not required during systemic infection.  相似文献   

19.
The microplasmodia of Physarum polycephalum express three types of β-glucosidases: secretory enzyme, a soluble cytoplasmic enzyme and a membrane-bound enzyme. We are interested in the physiological role of three enzymes. We report the sequence of cDNA for membrane β-glucosidase 1, which consists of 3825 nucleotides that includes an open reading frame encoding 1248 amino acids. The molecular weight of membrane β-glucosidase 1 was calculated to be 131,843 based on the predicted amino acid composition. Glycosyl hydrolase family 3 N-terminal and C-terminal domains were found within the N-terminal half of the membrane β-glucosidase 1 sequence and were highly homologous with the primary structures of fungal β-glucosidases. Notably, the C-terminal half of membrane β-glucosidase 1 contains two calx-β motifs, which are known to be Ca2+ binding domains in the Drosophila Na+/Ca2+ exchanger; an RGD sequence, which is known to be a cell attachment sequence; and a transmembrane region. In this way, Physarum membrane β-glucosidase 1 differs from all previously identified family 3 β-glucosidases. In addition to cDNA for membrane β-glucosidase 1, two other distinctly different mRNAs were also isolated. Two sequences were largely identical to cDNA for membrane β-glucosidase 1, but included a long insert sequence having a stop codon, leading to truncation of their products, which could account for other β-glucosidase forms occurred in Physarum poycephalum.

Thus, the membrane β-glucosidase is a new type family 3 enzyme fused with the Calx-β domain. We propose that Calx-β domain may modulate the β-glucosidase activity in response to changes in the Ca2+ concentration.  相似文献   


20.
M H O'Leary  R M Herreid 《Biochemistry》1978,17(6):1010-1014
Ornithine decarboxylase from Lactobacillus 30a is gradually inactivated by treatment with alpha-methylornithine, but activity is restored by treatment of the inactivated enzyme with pyridoxal phosphate. Inactivation of the enzyme is associated with formation of pyridoxamine phosphate and 5-amino-2-pentanone, alpha-Methylornithine is decarboxylated by the enzyme about 6000 times more slowly than is ornithine under the same conditions. These observations provide an explanation for the previously observed inhibition of ornithine decarboxylase by alpha-methylornithine [M. M. Adbel-Monem, N. E. Newton, and C. E. Weeks (1974), J. Med. Chem. 17, 4447]: alpha-Methylornithine undergoes a decarboxylation-dependent transamination as a result of incorrect protonation of the quinoid intermediate which is formed by decarboxylation of the enzyme-bound pyridoxal phosphate-substrate Schiff base. This protonation produces inactive enzyme. Decarboxylation of ornithine by this enzyme produces a small amount of 4-aminobutanal, presumably also by decarboxylation-dependent transamination.  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