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.
In this study, we report two high‐resolution structures of the pyridoxal 5′ phosphate (PLP)‐dependent enzyme kynurenine aminotransferase‐I (KAT‐I). One is the native structure with the cofactor in the PLP form bound to Lys247 with the highest resolution yet available for KAT‐I at 1.28 Å resolution, and the other with the general PLP‐dependent aminotransferase inhibitor, aminooxyacetate (AOAA) covalently bound to the cofactor at 1.54 Å. Only small conformational differences are observed in the vicinity of the aldimine (oxime) linkage with which the PLP forms the Schiff base with Lys247 in the 1.28 Å resolution native structure, in comparison to other native PLP‐bound structures. We also report the inhibition of KAT‐1 by AOAA and aminooxy‐phenylpropionic acid (AOPP), with IC50s of 13.1 and 5.7 μM, respectively. The crystal structure of the enzyme in complex with the inhibitor AOAA revealed that the cofactor is the PLP form with the external aldimine linkage. The location of this oxime with the PLP, which forms in place of the native internal aldimine linkage of PLP of the native KAT‐I, is away from the position of the native internal aldimine, with the free Lys247 substantially retaining the orientation of the native structure. Tyr101, at the active site, was observed in two conformations in both structures.  相似文献   

2.
Histamine is an important chemical mediator for a wide variety of physiological reactions. l-Histidine decarboxylase (HDC) is the primary enzyme responsible for histamine synthesis and produces histamine from histidine in a one-step reaction. In this study, we determined the crystal structure of human HDC (hHDC) complexed with the inhibitor histidine methyl ester. This structure shows the detailed features of the pyridoxal-5'-phosphate inhibitor adduct (external aldimine) at the active site of HDC. Moreover, a comparison of the structures of hHDC and aromatic l-amino acid (l-DOPA) decarboxylase showed that Ser-354 was a key residue for substrate specificity. The S354G mutation at the active site enlarged the size of the hHDC substrate-binding pocket and resulted in a decreased affinity for histidine, but an acquired ability to bind and act on l-DOPA as a substrate. These data provide insight into the molecular basis of substrate recognition among the group II pyridoxal-5'-phosphate-dependent decarboxylases.  相似文献   

3.
Pyridoxal 5′‐phosphate (PLP)‐dependent β‐transaminases (βTAs) reversibly catalyze transamination reactions by recognizing amino groups linked to the β‐carbon atoms of their substrates. Although several βTA structures have been determined as holo forms containing PLP, little is known about the effect of PLP on the conversion of the apo structure to the holo structure. We determined the crystal structure of the apo form of a βTA from Mesorhizobium sp. strain LUK at 2.2 Å resolution to elucidate how PLP affects the βTA structure. The structure revealed three major disordered regions near the active site. Structural comparison with the holo form also showed that the disordered regions in the apo form are ordered and partially adopt secondary structures in the holo form. These findings suggest that PLP incorporation into the active site contributes to the structural stability of the active site architecture, thereby forming the complete active site. Our results provide novel structural insights into the role of PLP in terms of active site formation.  相似文献   

4.
This report describes a microassay procedure for mammalian histidine decarboxylase (HDC) based on the measurement of [14C]O2 formed from l-[1-14C]histidine. This assay is particularly useful for quick measurement of HDC activity both in microgram quantities of cell or tissue extract and in tissues that contain significant levels of endogenous histamine.Using this assay, we have shown that the pH optimum, Km and thermolability of HDC are similar for extracts prepared both from normal rat peritoneal mast cells and from the Furth mouse mastocytoma. HDC activity could be detected in homogenates prepared from 105 rat mast cells, and it was expressed on a per cell basis. Mast cell HDC activity varied with the strain of rat from which the cells were obtained and with the season when they were assayed.  相似文献   

5.
Histidine decarboxylase (HDC) from Lactobacillus 30a converts histidine to histamine, a process that enables the bacteria to maintain the optimum pH range for cell growth. HDC is regulated by pH; it is active at low pH and inactive at neutral to alkaline pH. The X-ray structure of HDC at pH 8 revealed that a helix was disordered, resulting in the disruption of the substrate-binding site. The HDC trimer has also been shown to exhibit cooperative kinetics at neutral pH, that is, histidine can trigger a T-state to R-state transition. The D53,54N mutant of HDC has an elevated Km, even at low pH, indicating that the enzyme assumes the low activity T-state. We have solved the structures of the D53,54N mutant at low pH, with and without the substrate analog histidine methyl ester (HME) bound. Structural analysis shows that the apo-D53,54N mutant is in the inactive or T-state and that binding of the substrate analog induces the enzyme to adopt the active or R-state. A mechanism for the cooperative transition is proposed.  相似文献   

6.
Histidine decarboxylase (HDC) from Enterobacter aerogenes DL-1 was purified in a three-step procedure involving ammonium sulfate precipitation, Sephadex G-100, and DEAE-Sepharose column chromatography. The partially purified enzyme showed a single protein band of 52.4 kD on sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE). The optimum pH for HDC activity was 6.5, and the enzyme was stable between pH 4 and 8. Enterobacter aerogenes HDC had optimal activity at 40°C and retained most of its activity between 4 and 50°C. HDC activity was reduced in the presence of numerous tested compounds. Particularly with SDS, it significantly (p < 0.01) inhibited enzyme activity. Conversely, Ca2+ and Mn2+ showed prominent activation effects (p < 0.01) with activity increasing to 117.20% and 123.42%, respectively. The Lineweaver–Burk plot showed that K m and V max values of the enzyme for L-histidine were 0.21 mM and 71.39 µmol/min, respectively. In comparison with most HDCs from other microorganisms and animals, HDC from E. aerogenes DL-1 displayed higher affinity and greater reaction velocity toward L-histidine.  相似文献   

7.
8.
The concentration of mouse kidney histidine decarboxylase (HDC) is modulated by estrogen, testosterone, and thyroxine in a tissue-specific manner. Variation in HDC levels between strains of mice can be used to investigate the genetic regulation of (i) enzyme structure, (ii) tissue specific expression, and (iii) induction and repression by hormones. Variation in the structure of HDC between different inbred strains of mice affecting its K m for the cofactor pyridoxal-5-phosphate (PLP) and its heat stability has been discovered. The alternative phenotypes are additively inherited in crosses and the heat stability difference is due to alleles of a single structural gene, Hdc-s, which segregate among the BXD and BXH recombinant inbred strains. The allele Hdc-s b determines the heat-stable phenotype (C57BL substrains), and the allele Hdc-s d the heat-labile phenotype (DBA/2 and C3H/He strains). The alleles of the structural gene cosegregate with alleles of a regulatory gene previously named Hdc (determining kidney enzyme concentration); there were no recombinants among 38 RI strains. Therefore the two loci are less than 0.685 cM apart and comprise part of the HDC gene complex, [Hdc], on chromosome 2 of the mouse.This work was supported in part by an SERC studentship to S.A.M. and an MRC project grant to G.B.  相似文献   

9.
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   

10.
Histidine decarboxylase (HDC) is an enzyme that converts histidine to histamine. Inhibition of HDC has several medical applications, and HDC inhibitors are of considerable interest for the study of histidine metabolism. (S)-α-Fluoromethylhistidine di-hydrochloride (α-FMH) is a potent HDC inhibitor that is commercially available at high cost in small amounts only. Here we report a novel, inexpensive, and efficient method for synthesis of α-FMH using methyl 2-aziridinyl-3-(N-triphenylmethyl-4-imidazolyl) propionate and HF/pyridine, with experimental yield of 57%. To identify novel targets for α-FMH, we developed a three step in silico work-flow for identifying physically plausible protein targets. The work-flow resulted in 21 protein target hits, including several enzymes involved in glutathione metabolism, and notably, two isozymes of the glutathione S-transferase (GST) superfamily, which plays a central role in drug metabolism. In view of this predictive data, the efficacy of α-FMH as a GST inhibitor was investigated in vitro. α-FMH was demonstrated to be an effective inhibitor of GST at micromolar concentration, suggesting that off-target effects of α-FMH may limit physiological drug metabolism and elimination by GST-dependent mechanisms. The present study therefore provides new avenues for obtaining α-FMH and for studying its biochemical effects, with potential implications for drug development.  相似文献   

11.
Histamine (HA) is present in substantial quantities in all ganglia of Aplysia californica. Within the cerebral ganglia this amine is known to be concentrated in at least two identified neurons designated C-2 neurons. In this study a combination of chemical and enzymatic analyses was employed to provide evidence for the existence of a biochemical pathway for HA synthesis in ganglia and individual neurons of this marine mollusk. Examination of extracts of individual neurons dissected from ganglia organ-cultured in the presence of [3H]histidine showed that every neuron accumulated labelled histidine, but only the HA-containing C-2 neurons synthesized and stored labelled HA suggesting that the formation of HA in Aplysia could be catalyzed by the enzyme histidine decarboxylase (HDC). HDC activity was studied with a new microradiometric assay. Many of the properties of the molluscan HDC studied were found to correspond to the vertebrate enzyme. Enzyme activity was inhibited by α-hydrazino-histidine but unaffected by concentrations of α-methyldopa or by 5-(3,4-dihydroxycinnamoyl) salicylic acid which produced nearly complete inhibition of aromatic amino acid decarboxylase activity. HDC was measurable in nervous but not other Aplysia tissues assayed. All 5 major ganglia contained HDC activity which spanned a 15-fold range between the least and most active ganglia. Only 4 of the 13 nerve trunks assayed yielded measurable enzymic activity; these active nerves were associated with the cerebral ganglia which has the highest HDC activity of all measured ganglia. Of the numerous individual neurons assayed for HDC, only the C-2 cells showed measurable enzyme activity, about 25 pmol/cell/h or 70 μmol/g protein/h. Since the activity of HDC in the HA-containing neurons was at least three orders of magnitude larger than all other neurons assayed in the cerebral and other ganglia, these data appear to provide a direct metabolic basis for the selective presence of HA in these cells, and they indicate that the cellular presence of HDC provides a useful biochemical marker for the location of HA-rich neurons in Aplysia.  相似文献   

12.
In this study, HPLC-MS and ICP-MS methods were used for the determination of histidine and nickel in Morus L., Robinia pseudoacacia L., and Populus nigra L. leaves taken from industrial areas including Gaziantep and Bursa cities. In the determination of histidine by HPLC-MS, all of the system parameters such as flow rate of mobile phase, fragmentor potential, injection volume and column temperature were optimized and found to be 0.2 mL min?1, 70 V, 15 µL, and 20°C, respectively. Under the optimum conditions, histidine was extracted from plant sample by distilled water at 90°C for 30 min. Concentrations of histidine as mg kg?1 were found to be between 2–9 for Morus L., 6–13 for Robinia pseudoacacia L., and 2–10 for Populus nigra L. Concentrations of nickel were in the ranges of 5–10 mg kg?1 for Morus L., 3–10 mg kg?1 for Robinia pseudoacacia L., and 0.6–4 mg kg?1 for Populus nigra L. A significant linear correlation (r = 0.78) between histidine and Ni was observed for Populus nigra L., whereas insignificant linear correlation for Robinia pseudoacacia L. (r = 0.22) were seen. Limits of detection (LOD) and quantitation (LOQ) were found to be 0.025 mg Ni L?1 and 0.075 mg Ni L?1, respectively.  相似文献   

13.
Adenosine triphosphate phosphoribosyltransferase (ATP‐PRT) catalyzes the first committed step of the histidine biosynthesis in plants and microorganisms. Here, we present the functional and structural characterization of the ATP‐PRT from the pathogenic ε‐proteobacteria Campylobacter jejuni (CjeATP‐PRT). This enzyme is a member of the long form (HisGL) ATP‐PRT and is allosterically inhibited by histidine, which binds to a remote regulatory domain, and competitively inhibited by AMP. In the crystalline form, CjeATP‐PRT was found to adopt two distinctly different hexameric conformations, with an open homohexameric structure observed in the presence of substrate ATP, and a more compact closed form present when inhibitor histidine is bound. CjeATP‐PRT was observed to adopt only a hexameric quaternary structure in solution, contradicting previous hypotheses favoring an allosteric mechanism driven by an oligomer equilibrium. Instead, this study supports the conclusion that the ATP‐PRT long form hexamer is the active species; the tightening of this structure in response to remote histidine binding results in an inhibited enzyme.  相似文献   

14.
The pyridoxal 5‐phosphate (PLP) cofactor is a significant organic molecule in medicinal chemistry. It is often found covalently bound to lysine residues in proteins to form PLP dependent enzymes. An example of this family of PLP dependent enzymes is γ‐aminobutyric acid aminotransferase (GABA‐AT) which is responsible for the degradation of the neurotransmitter GABA. Its inhibition or inactivation can be used to prevent the reduction of GABA concentration in brain which is the source of several neurological disorders. As a test case for PLP dependent enzymes, we have performed molecular dynamics simulations of GABA‐AT to reveal the roles of the protein residues and its cofactor. Three different states have been considered: the apoenzyme, the holoenzyme, and the inactive state obtained after the suicide inhibition by vigabatrin. Different protonation states have also been considered for PLP and two key active site residues: Asp298 and His190. Together, 24 independent molecular dynamics trajectories have been simulated for a cumulative total of 2.88 µs. Our results indicate that, unlike in aqueous solution, the PLP pyridine moiety is protonated in GABA‐AT. This is a consequence of a pKa shift triggered by a strong charge–charge interaction with an ionic “diad” formed by Asp298 and His190 that would help the activation of the first half‐reaction of the catalytic mechanism in GABA‐AT: the conversion of PLP to free pyridoxamine phosphate (PMP). In addition, our MD simulations exhibit additional strong hydrogen bond networks between the protein and PLP: the phosphate group is held in place by the donation of at least three hydrogen bonds while the carbonyl oxygen of the pyridine ring interacts with Gln301; Phe181 forms a π–π stacking interaction with the pyridine ring and works as a gate keeper with the assistance of Val300. All these interactions are hypothesized to help maintain free PMP in place inside the protein active site to facilitate the second half‐reaction in GABA‐AT: the regeneration of PLP‐bound GABA‐AT (i.e., the holoenzyme). Proteins 2016; 84:875–891. © 2016 Wiley Periodicals, Inc.  相似文献   

15.
Alar, a Pyridoxal 5′-phosphate (PLP)-dependent bacterial enzyme is responsible for the racemisation of L-alanine into D-alanine which is essential for the peptidoglycan biosynthesis in both Gram-positive and Gram-negative bacteria. In the present study, we performed induced fit docking, binding free energy calculation and molecular dynamics simulation to elucidate the Alar inhibition potential of 1,2,4-thiadiazolidine-3,5-dione-based inhibitor 1. The inhibitor binds to the hydrophobic groove of Alar and the binding was found to be stable throughout 20-ns MD simulation. Induced fit docking result showed that Lys42, Tyr46, Tyr175 and Tyr364 residues are primarily responsible for the stabilisation of inhibitor–protein complex. Further, high negative van der Waals binding free energy value of –38.88 kcal/mol, indicated it as the main driving force for the inhibitor binding. Based on the information obtained from this study, we designed few molecules as potent Alar inhibitor. In order to gain structural insight and to validate the stability of complex, we performed 20-ns MD simulation of the designed molecule D1. Results obtained from this study can be used for the design of M. tuberculosis Alar potent inhibitors lacking affinity for the co-factor PLP.  相似文献   

16.
Abstract

An experiment was conducted in young pigs (initial BW 10.1 kg) to estimate the maintenance requirement for histidine and its efficiency of utilization for protein accretion using a comparative slaughter technique. Three groups of six pigs each were fed a purified diet supplying 0, 14 or 56 mg histidine per kg BW0.75. Following 21 d of feeding, pigs were killed for whole body compositional analysis. A representative group of six pigs was killed at the beginning of the experiment. Retention of histidine and total N were the main criteria of response. Histidine retention (R2 = 0.73) and N retention (R2 = 0.78) were linear functions of histidine intake (p < 0.001). Histidine requirement for zero histidine retention was 15.5 mg/kg BW0.75, whereas histidine required for zero N retention was 4.1 mg/kg BW0.75. At zero histidine retention, the pigs retained daily 82 mg N/kg BW0.75, presumably due to the degradation of histidine-rich compounds such as haemoglobin and/or carnosine. The slope of the regression line relating histidine retention to N retention indicated that 105 mg of histidine was deposited per gram of total N which was considerably less than the estimated histidine concentration in body protein (179 mg/g N). Based on the slopes of regression equations for histidine and N retention, marginal efficiency of histidine utilization was calculated to be 0.94 and 1.34, respectively.  相似文献   

17.
2-Phenylethanol (2-PE) is a desirable compound in the food and perfumery industries with a characteristic rose fragrance. Until now, most of the studied biotechnological processes to produce 2-PE were conducted using natural 2-PE-producing yeasts. Only several researches were conducted in other genetically engineered microorganisms that simulated the Ehrlich pathway for the conversion of amino acids to fusel alcohols. Here, a novel metabolic pathway has been designed in Escherichia coli to produce 2-PE, using the Rosa hybrid phenylacetaldehyde synthase (PAAS), a pyridoxal 5′-phosphate (PLP)-dependent enzyme capable of transforming l-phenylalanine (l-phe) into phenylacetaldehyde by decarboxylation and oxidation. To overcome the enzyme insolubility in E. coli, several plasmids and host strains were tested for their expression ability. The desired results were obtained by using the pTYB21 plasmid containing the intein tag from the Saccharomyces cerevisiae VMA1. It was discovered that the intein PAAS activity is temperature-dependent, working well in the range of 25 to 30 °C but losing most of its activity at 37 °C. When external PLP cofactor was added, the cells produced 0.39 g l-1 2-PE directly from l-phe. In addition, a biotransformation that was based only on internal de novo PLP synthesis produced 0.34 g l-1 2-PE, thus creating for the first time an E. coli strain that can produce 2-PE from l-phe without the need for exterior cofactor additions.  相似文献   

18.
Anatoxin-a(s) is a guanidine methyl phosphate ester (unprotonated molecular ion equals 252 daltons) isolated from the freshwater cyanobac-terium (blue-green alga) Anabaena flos-aquae strain NRC 525–17. Previous work has shown anatoxin-a(s) to be a potent irreversible inhibitor of electric eel ace-tylcholinesterase (EC 3.1.1.7, AChE). In the present study the interaction of anatoxin-a(s) with AChE was investigated by protection studies and since similarities have been noted between anatoxin-a(s) and the synthetic organophosphate anticholinesterases, the ability of reactivators to reactivate the inhibited enzyme was investigated. Treatments directed toward eliminating poisoning symptoms and in vivo protection from anatoxin-a(s) poisonings were investigated using oxime reactivators and atropine or pretreatment with a carbamate and atropine. Anatoxin-a(s) was shown to be an active site-directed inhibitor of acetyl-cholinesterase which is resistant to oxime reactivation due to the structure of its enzyme adduct. In vivo pretreatment with physostigmine and high concentrations of 2-PAM were the only effective antagonists against a lethal dose of anatoxin-a(s).  相似文献   

19.
To clarify the base of in vivo biological activities of peptidoglycans of Gram-positive bacteria, the effects of a polysaccharide peptide of Staphylococcus epidermidis peptidoglycan (SEPS) on the synthesis of histamine and putrescine in BALB/c mice were examined and compared with those of a lipopolysaccharide (LPS or endotoxin) of Gram-negative bacteria. Within a few hours after its injection into BALB/c mice, SEPS induced histidine decarboxylase (HDC), the enzyme forming histamine, in the liver, lung, spleen and bone marrow, and ornithine decarboxylase (ODC), the enzyme forming putrescine, in the tissues except for the lung. SEPS induced HDC activity even in mast cell-deficient mice and in nude mice. These effects of SEPS were essentially the same as those of LPS. However, the dosage of SEPS capable of inducing HDC and ODC was much higher (100 to 1,000 times) than that of LPS. We have reported that C3H/HeN mice are resistant to SEPS in producing acute arthritis, and their productions of IL-1 and prostaglandin E2 are less than BALB/c mice sensitive to producing acute arthritis. In the present study, it was also found that C3H/HeN mice were markedly resistant to SEPS in inducing HDC activity.  相似文献   

20.

Background

O-acetyl serine sulfhydrylase (OASS) is a pyridoxal phosphate (PLP) dependent enzyme catalyzing the last step of the cysteine biosynthetic pathway. Here we analyze and investigate the factors responsible for recognition and different conformational changes accompanying the binding of various ligands to OASS.

Methods

X ray crystallography was used to determine the structures of OASS from Entamoeba histolytica in complex with methionine (substrate analog), isoleucine (inhibitor) and an inhibitory tetra-peptide to 2.00 Å, 2.03 Å and 1.87 Å resolutions, respectively. Molecular dynamics simulations were used to investigate the reasons responsible for the extent of domain movement and cleft closure of the enzyme in presence of different ligands.

Results

Here we report for the first time an OASS-methionine structure with an unmutated catalytic lysine at the active site. This is also the first OASS structure with a closed active site lacking external aldimine formation. The OASS-isoleucine structure shows the active site cleft in open state. Molecular dynamics studies indicate that cofactor PLP, N88 and G192 form a triad of energy contributors to close the active site upon ligand binding and orientation of the Schiff base forming nitrogen of the ligand is critical for this interaction.

Conclusions

Methionine proves to be a better binder to OASS than isoleucine. The β branching of isoleucine does not allow it to reorient itself in suitable conformation near PLP to cause active site closure.

General significance

Our findings have important implications in designing better inhibitors against OASS across all pathogenic microbial species.  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