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.
2-Methyl-3-hydroxypyridine-5-carboxylic acid (MHPC) oxygenase (MHPCO) is a flavoprotein that catalyzes the oxygenation of MHPC to form alpha-(N-acetylaminomethylene)-succinic acid. Although formally similar to the oxygenation reactions catalyzed by phenol hydroxylases, MHPCO catalyzes the oxygenation of a pyridyl derivative rather than a simple phenol. Therefore, in this study, the mechanism of the reaction was investigated by replacing the natural cofactor FAD with FAD analogues having various substituents (-Cl, -CN, -NH(2), -OCH(3)) at the C8-position of the isoalloxazine. Thermodynamic and catalytic properties of the reconstituted enzyme were investigated and found to be similar to those of the native enzyme, validating that these FAD analogues are reasonable to be used as mechanistic probes. Dissociation constants for the binding of MHPC or the substrate analogue 5-hydroxynicotinate (5HN) to the reconstituted enzymes indicate that the reconstituted enzymes bind well with ligands. Redox potential values of the reconstituted enzymes were measured and found to be more positive than the values of free FAD analogues, which correlated well with the electronic effects of the 8-substituents. Studies of the reductive half-reaction of MHPCO have shown that the rates of flavin reduction by NADH could be described as a parabolic relationship with the redox potential values of the reconstituted enzymes, which is consistent with the Marcus electron transfer theory. Studies of the oxidative half-reaction of MHPCO revealed that the rate of hydroxylation depended upon the different analogues employed. The rate constants for the hydroxylation step correlated with the calculated pK(a) values of the 8-substituted C(4a)-hydroxyflavin intermediates, which are the leaving groups in the oxygen transfer step. It was observed that the rates of hydroxylation were greater when the pK(a) values of C(4a)-hydroxyflavins were lower. Although these results are not as dramatic as those from analogous studies with parahydroxybenzoate hydroxylase (Ortiz-Maldonado et al., (1999) Biochemistry 38, 8124-8137), they are consistent with the model that the oxygenation reaction of MHPCO occurs via an electrophilic aromatic substitution mechanism analogous to the mechanisms for parahydroxybenzoate and phenol hydroxylases.  相似文献   

2.
Many flavoenzymes catalyze hydroxylation of aromatic compounds especially phenolic compounds have been isolated and characterized. These enzymes can be classified as either single‐component or two‐component flavin‐dependent hydroxylases (monooxygenases). The hydroxylation reactions catalyzed by the enzymes in this group are useful for modifying the biological properties of phenolic compounds. This review aims to provide an in‐depth discussion of the current mechanistic understanding of representative flavin‐dependent monooxygenases including 3‐hydroxy‐benzoate 4‐hydroxylase (PHBH, a single‐component hydroxylase), 3‐hydroxyphenylacetate 4‐hydroxylase (HPAH, a two‐component hydroxylase), and other monooxygenases which catalyze reactions in addition to hydroxylation, including 2‐methyl‐3‐hydroxypyridine‐5‐carboxylate oxygenase (MHPCO, a single‐component enzyme that catalyzes aromatic‐ring cleavage), and HadA monooxygenase (a two‐component enzyme that catalyzes additional group elimination reaction). These enzymes have different unique structural features which dictate their reactivity toward various substrates and influence their ability to stabilize flavin intermediates such as C4a‐hydroperoxyflavin. Understanding the key catalytic residues and the active site environments important for governing enzyme reactivity will undoubtedly facilitate future work in enzyme engineering or enzyme redesign for the development of biocatalytic methods for the synthesis of valuable compounds.  相似文献   

3.
4.
The amino acid ligands to the active site iron in the aromatic amino acid hydroxylase tyrosine hydroxylase are two histidines and a glutamate. This 2-histidine-1-carboxylate motif has been found in a number of other metalloenzymes which catalyze a variety of oxygenase reactions. As a probe of the plasticity of this metal binding site, each of the ligands in TyrH has been mutated to glutamine, glutamate, or histidine. The H336E and H336Q enzymes show dramatic decreases in iron affinity but retain substantial activity for both tyrosine hydroxylation and tetrahydropterin oxidation. The H331E enzyme shows a lesser decrease in iron affinity and is unable to hydroxylate tyrosine. Instead, this enzyme oxidizes tetrahydropterin in the absence of added tyrosine. The E376H enzyme has no significant activity, while the E376Q enzyme hydroxylates tyrosine at about 0.4% the wild-type rate. When dopamine is bound to either the H336Q or H331E enzymes, the position of the long wavelength charge-transfer absorbance band is consistent with the change in the metal ligand. In contrast, the H336E enzyme does not form a stable binary complex with dopamine, while the E376H and E376Q enzymes catalyze dopamine oxidation.  相似文献   

5.
W H Ward  A R Fersht 《Biochemistry》1988,27(15):5525-5530
Tyrosyl-tRNA synthetase from Bacillus stearothermophilus is a classical example of an enzyme with half-of-the-sites activity. The enzyme crystallizes as a symmetrical dimer that is composed of identical subunits, each having a complete active site. In solution, however, tyrosyl-tRNA synthetase binds tightly, and activates rapidly, only 1 mol of Tyr/mol of dimer. It has recently been shown that the half-of-the-sites activity results from an inherent asymmetry of the enzyme. Only one subunit catalyzes formation of Tyr-AMP, and interchange of activity between subunits is not detectable over a long time scale. Paradoxically, however, the kinetics of tRNA charging are biphasic with respect to [Tyr], suggesting that both subunits of the dimer are catalytically active. This paradox has now been resolved by kinetic analysis of heterodimeric enzymes containing different mutations in each subunit. Biphasic kinetics with unchanged values of KM for Tyr are maintained when one of the two tRNA-binding domains is removed and also when the affinity of the "inactive" site for Try is reduced by 2-58-fold. The biphasic kinetics do not result from catalysis at both active sites, but instead appear to result from two molecules of Tyr binding sequentially to the same site. A second molecule of Tyr perhaps aids the dissociation of Tyr-tRNA by displacing the tyrosyl moiety from its binding site. A monomer of the enzyme is probably too small to allow both recognition and aminoacylation of a tRNA molecule. This could explain the requirement for the enzyme to function as an asymmetric dimer.  相似文献   

6.
The compound 4-(fluorosulfonyl)benzoic acid (4-FSB) functions as an affinity label of the dimeric pig lung pi class glutathione S-transferase yielding a completely inactive enzyme. Protection against inactivation is provided by glutathione-based ligands, suggesting that the reaction target is near or part of the glutathione binding site. Radioactive 4-FSB is incorporated to the extent of 1 mol per mole of enzyme subunit. Peptide mapping revealed that 4-FSB reacts with two tyrosine residues in the ratio 69% Tyr7 and 31% Tyr106. The ratio is not changed by the addition of ligands. The results suggest that only one of the tyrosine residues can be labeled in the active site of a given subunit; i.e., reactions with Tyr7 and Tyr106 are mutually exclusive. We propose that the difference in labeling of these tyrosine residues is related to their pKa values, with Tyr7 exhibiting the lower pKa. The modified enzyme no longer binds to a S-hexylglutathione-agarose affinity column, even when only one of the active sites contains 4-FSB; these results may reflect interaction between the subunits. We conclude that Tyr7 and Tyr106 of the pig lung class pi glutathione S-transferase are important for function and are located at or close to the substrate binding site of the enzyme.  相似文献   

7.
The crystal structure of the 3-chlorocatechol 1,2-dioxygenase from the Gram-positive bacterium Rhodococcus opacus (erythropolis) 1CP, a Fe(III) ion-containing enzyme specialized in the aerobic biodegradation of 3-chloro- and methyl-substituted catechols, has been solved by molecular replacement techniques using the coordinates of 4-chlorocatechol 1,2-dioxygenase from the same organism (PDB code 1S9A) as a starting model and refined at 1.9 A resolution (R(free) 21.9%; R-factor 17.4%). The analysis of the structure and of the kinetic parameters for a series of different substrates, and the comparison with the corresponding data for the 4-chlorocatechol 1,2-dioxygenase isolated from the same bacterial strain, provides evidence of which active site residues are responsible for the observed differences in substrate specificity. Among the amino acid residues expected to interact with substrates, only three are altered Val53(Ala53), Tyr78(Phe78) and Ala221(Cys224) (3-chlorocatechol 1,2-dioxygenase(4-chlorocatechol 1,2-dioxygenase)), clearly identifying the substitutions influencing substrate selectivity in these enzymes. The crystallographic asymmetric unit contains eight subunits (corresponding to four dimers) that show heterogeneity in the conformation of a co-crystallized molecule bound to the catalytic non-heme iron(III) ion resembling a benzohydroxamate moiety, probably a result of the breakdown of recently discovered siderophores synthesized by Gram-positive bacteria. Several different modes of binding benzohydroxamate into the active site induce distinct conformations of the interacting protein ligands Tyr167 and Arg188, illustrating the plasticity of the active site origin of the more promiscuous substrate preferences of the present enzyme.  相似文献   

8.
The 3D structure of the flavoprotein D-amino acid oxidase (DAAO) from the yeast Rhodotorula gracilis (RgDAAO) in complex with the competitive inhibitor anthranilate was solved (resolution 1.9A) and structural features relevant for the overall conformation and for catalytic activity are described. The FAD is bound in an elongated conformation in the core of the enzyme. Two anthranilate molecules are found within the active site cavity; one is located in a funnel forming the entrance, and the second is in contact with the flavin. The anchoring of the ligand carboxylate with Arg285 and Tyr223 is found for all complexes studied. However, while the active site group Tyr238-OH interacts with the carboxylate in the case of the substrate D-alanine, of D-CF(3)-alanine, or of L-lactate, in the anthranilate complex the phenol group rotates around the C2-C3 bond thus opening the entrance of the active site, and interacts there with the second bound anthranilate. This movement serves in channeling substrate to the bottom of the active site, the locus of chemical catalysis. The absence in RgDAAO of the "lid" covering the active site, as found in mammalian DAAO, is interpreted as being at the origin of the differences in kinetic mechanism between the two enzymes. This lid has been proposed to regulate product dissociation in the latter, while the side-chain of Tyr238 might exert a similar role in RgDAAO. The more open active site architecture of RgDAAO is the origin of its much broader substrate specificity. The RgDAAO enzyme forms a homodimer with C2 symmetry that is different from that reported for mammalian D-amino acid oxidase. This different mode of aggregation probably causes the differences in stability and tightness of FAD cofactor binding between the DAAOs from different sources.  相似文献   

9.
In the biosynthesis of several anthracyclines, aromatic polyketides produced by many Streptomyces species, the aglycone core is modified by a specific flavin adenine dinucleotide (FAD)- and NAD(P)H-dependent aklavinone-11-hydroxylase. Here, we report the crystal structure of a ternary complex of this enzyme from Streptomyces purpurascens, RdmE, with FAD and the substrate aklavinone. The enzyme is built up of three domains, a FAD-binding domain, a domain involved in substrate binding, and a C-terminal thioredoxin-like domain of unknown function. RdmE exhibits structural similarity to aromatic hydroxylases from the p-hydroxybenzoate hydroxylase family, but unlike most other related enzymes, RdmE is a monomer. The substrate is bound in a hydrophobic pocket in the interior of the enzyme, and access to this pocket is provided through a different route than for the isoalloxazine ring of FAD—the backside of the ligand binding cleft. The architecture of the substrate binding pocket and the observed enzyme-aklavinone interactions provide a structural explanation for the specificity of the enzyme for non-glycosylated substrates with C9-R stereochemistry. The isoalloxazine ring of the flavin cofactor is bound in the “out” conformation but can be modeled in the “in” conformation without invoking large conformational changes of the enzyme. This model places the flavin ring in a position suitable for catalysis, almost perpendicular to the tetracyclic ring system of the substrate and with a distance of the C4a carbon atom of the isoalloxazine ring to the C-11 carbon atom of the substrate of 4.8 Å. The structure suggested that a Tyr224-Arg373 pair might be involved in proton abstraction at the C-6 hydroxyl group, thereby increasing the nucleophilicity of the aromatic ring system and facilitating electrophilic attack by the perhydroxy-flavin intermediate. Replacement of Tyr224 by phenylalanine results in inactive enzyme, whereas mutants at position Arg373 retain catalytic activity close to wild-type level. These data establish an essential role of residue Tyr224 in catalysis, possibly in aligning the substrate in a position suitable for catalysis.  相似文献   

10.
3,4‐Dihydroxy‐2‐butanone‐4‐phosphate synthase (DHBPS) encoded by ribB gene is one of the first enzymes in riboflavin biosynthesis pathway and catalyzes the conversion of ribulose‐5‐phosphate (Ru5P) to 3,4‐dihydroxy‐2‐butanone‐4‐phosphate and formate. DHBPS is an attractive target for developing anti‐bacterial drugs as this enzyme is essential for pathogens, but absent in humans. The recombinant DHBPS enzyme of Salmonella requires magnesium ion for its activity and catalyzes the formation of 3,4‐dihydroxy‐2‐butanone‐4‐phosphate from Ru5P at a rate of 199 nmol min?1 mg?1 with Km value of 116 μM at 37°C. Further, we have determined the crystal structures of Salmonella DHBPS in complex with sulfate, Ru5P and sulfate‐zinc ion at a resolution of 2.80, 2.52, and 1.86 Å, respectively. Analysis of these crystal structures reveals that the acidic loop (residues 34–39) responsible for the acid‐base catalysis is disordered in the absence of substrate or metal ion at the active site. Upon binding either substrate or sulfate and metal ions, the acidic loop becomes stabilized, adopts a closed conformation and interacts with the substrate. Our structure for the first time reveals that binding of substrate Ru5P alone is sufficient for the stabilization of the acidic active site loop into a closed conformation. In addition, the Glu38 residue from the acidic active site loop undergoes a conformational change upon Ru5P binding, which helps in positioning the second metal ion that stabilizes the Ru5P and the reaction intermediates. This is the first structural report of DHBPS in complex with either substrate or metal ion from any eubacteria. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

11.
Two high-resolution structures of a double mutant of bacterial cholesterol oxidase in the presence or absence of a ligand, glycerol, are presented, showing the trajectory of glycerol as it binds in a Michaelis complex-like position in the active site. A group of three aromatic residues forces the oxidized isoalloxazine moiety to bend along the N5-N10 axis as a response to the binding of glycerol in the active site. Movement of these aromatic residues is only observed in the glycerol-bound structure, indicating that some tuning of the FAD redox potential is caused by the formation of the Michaelis complex during regular catalysis. This structural study suggests a possible mechanism of substrate-assisted flavin activation, improves our understanding of the interplay between the enzyme, its flavin cofactor and its substrate, and is of use to the future design of effective cholesterol oxidase inhibitors.  相似文献   

12.
Misquitta SA  Colman RF 《Biochemistry》2005,44(24):8608-8619
To study the communication between the two active sites of dimeric glutathione S-transferase A1-1, we used heterodimers containing one wild-type (WT) active site and one active site with a single mutation at either Tyr9, Arg15, or Arg131. Tyr9 and Arg15 are part of the active site of the same subunit, while Arg131 contributes to the active site of the opposite subunit. The V(max) values of Tyr9 and Arg15 mutant enzymes were less than 2% that of WT, indicating their importance in catalysis. In contrast, V(max) values of Arg131 mutant enzymes were about 50-90% of that of WT enzyme while K(m)(GSH) values were approximately 3-8 times that of WT, suggesting that Arg131 plays a role in glutathione binding. The mutant enzyme (with a His(6) tag) and the WT enzyme (without a His(6) tag) were used to construct heterodimers (WT-Y9F, WT-Y9T, WT-R15Q, WT-R131M, WT-R131Q, and WT-R131E) by incubation of a mixture of wild-type and mutant enzyme at pH 7.5 in buffer containing 1,6-hexanediol, followed by dialysis against buffer lacking the organic solvent. The resultant heterodimers were separated from the wild-type and mutant homodimers using chromatography on nickel-nitrilotriacetic acid agarose. The V(max) values of all heterodimers were lower than expected for independent active sites. Our experiments demonstrate that mutation of an amino acid residue in one active site affects the activity in the other active site. Modeling studies show that key amino acid residues and water molecules connect the two active sites. This connectivity is responsible for the cross-talk between the active sites.  相似文献   

13.
Some starch-degrading enzymes accommodate carbohydrates at sites situated at a certain distance from the active site. In the crystal structure of barley alpha-amylase 1, oligosaccharide is thus bound to the 'sugar tongs' site. This site on the non-catalytic domain C in the C-terminal part of the molecule contains a key residue, Tyr380, which has numerous contacts with the oligosaccharide. The mutant enzymes Y380A and Y380M failed to bind to beta-cyclodextrin-Sepharose, a starch-mimic resin used for alpha-amylase affinity purification. The K(d) for beta-cyclodextrin binding to Y380A and Y380M was 1.4 mm compared to 0.20-0.25 mm for the wild-type, S378P and S378T enzymes. The substitution in the S378P enzyme mimics Pro376 in the barley alpha-amylase 2 isozyme, which in spite of its conserved Tyr378 did not bind oligosaccharide at the 'sugar tongs' in the structure. Crystal structures of both wild-type and S378P enzymes, but not the Y380A enzyme, showed binding of the pseudotetrasaccharide acarbose at the 'sugar tongs' site. The 'sugar tongs' site also contributed importantly to the adsorption to starch granules, as Kd = 0.47 mg.mL(-1) for the wild-type enzyme increased to 5.9 mg.mL(-1) for Y380A, which moreover catalyzed the release of soluble oligosaccharides from starch granules with only 10% of the wild-type activity. beta-cyclodextrin both inhibited binding to and suppressed activity on starch granules for wild-type and S378P enzymes, but did not affect these properties of Y380A, reflecting the functional role of Tyr380. In addition, the Y380A enzyme hydrolyzed amylose with reduced multiple attack, emphasizing that the 'sugar tongs' participates in multivalent binding of polysaccharide substrates.  相似文献   

14.
AP endonuclease (AP endo), a key enzyme in repair of abasic sites in DNA, makes a single nick 5' to the phosphodeoxyribose of an abasic site (AP-site). We recently proposed a novel mechanism, whereby the enzyme uses a key tyrosine (Tyr(171)) to directly attack the scissile phosphate of the AP-site. We showed that loss of the tyrosyl hydroxyl from Tyr(171) resulted in dramatic diminution in enzymatic efficiency. Here we extend the previous work to compare binding/recognition of AP endo to oligomeric DNA with and without an AP-site by wild type enzyme and several tyrosine mutants including Tyr(128), Tyr(171) and Tyr(269). We used single turnover and electrophoretic mobility shift assays. As expected, binding to DNA with an AP-site is more efficient than binding to DNA without one. Unlike catalytic cleavage by AP endo, which requires both hydroxyl and aromatic moieties of Tyr(171), the ability to bind DNA efficiently without an AP-site is independent of an aromatic moiety at position 171. However, the ability to discriminate efficiently between DNA with and without an AP-site requires tyrosine at position 171. Thus, AP endo requires a tyrosine at the active site for the properties that enable it to behave as an efficient, processive endonuclease.  相似文献   

15.
Rieske oxygenase (RO) systems are two- and three-component enzyme systems that catalyze the formation of cis-dihydrodiols from aromatic substrates. Degradation of pollutants in contaminated soil and generation of chiral synthons have been the major foci of RO research. Substrate specificity and product regio- and stereoselectivity have been shown to vary between individual ROs. While directed evolution methods for altering RO function have been successful in the past, rational engineering of these enzymes still poses a challenge due to the lack of structural understanding. Here we examine the structural changes induced by mutation of Phe-352 in naphthalene 1,2-dioxygenase from Pseudomonas sp. strain NCIB 9816-4 (NDO-O(9816-4)). Structures of the Phe-352-Val mutant in native form and in complex with phenanthrene and anthracene, along with those of wild-type NDO-O(9816-4) in complex with phenanthrene, anthracene, and 3-nitrotoluene, are presented. Phenanthrene was shown to bind in a different orientation in the Phe-352-Val mutant active site from that in the wild type, while anthracene was found to bind in similar positions in both enzymes. Two orientations of 3-nitrotoluene were observed, i.e., a productive and a nonproductive orientation. These orientations help explain why NDO-O(9816-4) forms different products from 3-nitrotoluene than those made from nitrobenzene dioxygenase. Comparison of these structures among themselves and with other known ROs bound to substrates reveals that the orientation of substrate binding at the active site is the primary determinant of product regio- and stereoselectivity.  相似文献   

16.
The three-dimensional structures of brain pyridoxal kinase and its complex with the nucleotide ATP have been elucidated in the dimeric form at 2.1 and 2.6 A, respectively. Results have shown that pyridoxal kinase, as an enzyme obeying random sequential kinetics in catalysis, does not possess a lid shape structure common to all kinases in the ribokinase superfamily. This finding has been shown to be in line with the condition that pyridoxal kinase binds substrates with variable sizes of chemical groups at position 4 of vitamin B(6) and its derivatives. In addition, the enzyme contains a 12-residue peptide loop in the active site for the prevention of premature hydrolysis of ATP. Conserved amino acid residues Asp(118) and Tyr(127) in the peptide loop could be moved to a position covering the nucleotide after its binding so that its chance to hydrolyze in the aqueous environment of the active site was reduced. With respect to the evolutionary trend of kinase enzymes, the existence of this loop in pyridoxal kinase could be classified as an independent category in the ribokinase superfamily according to the structural feature found and mechanism followed in catalysis.  相似文献   

17.
Reaction of rat liver glutathione S-transferase, isozyme 1-1, with 4-(fluorosulfonyl)benzoic acid (4-FSB), a xenobiotic substrate analogue, results in a time-dependent inactivation of the enzyme to a final value of 35% of its original activity when assayed at pH 6.5 with 1-chloro-2,4-dinitrobenzene (CDNB) as substrate. The rate of inactivation exhibits a nonlinear dependence on the concentration of 4-FSB from 0.25 mM to 9 mM, characterized by a KI of 0.78 mM and kmax of 0.011 min-1. S-Hexylglutathione or the xenobiotic substrate analogue, 2,4-dinitrophenol, protects against inactivation of the enzyme by 4-FSB, whereas S-methylglutathione has little effect on the reaction. These experiments indicate that reaction occurs within the active site of the enzyme, probably in the binding site of the xenobiotic substrate, close to the glutathione binding site. Incorporation of [3,5-3H]-4-FSB into the enzyme in the absence and presence of S-hexylglutathione suggests that modification of one residue is responsible for the partial loss of enzyme activity. Tyr 8 and Cys 17 are shown to be the reaction targets of 4-FSB, but only Tyr 8 is protected against 4-FSB by S-hexylglutathione. DTT regenerates cysteine from the reaction product of cysteine and 4-FSB, but does not reactivate the enzyme. These results show that modification of Tyr 8 by 4-FSB causes the partial inactivation of the enzyme. The Michaelis constants for various substrates are not changed by the modification of the enzyme. The pH dependence of the enzyme-catalyzed reaction of glutathione with CDNB for the modified enzyme, as compared with the native enzyme, reveals an increase of about 0.9 in the apparent pKa, which has been interpreted as representing the ionization of enzyme-bound glutathione; however, this pKa of about 7.4 for modified enzyme remains far below the pK of 9.1 for the -SH of free glutathione. Previously, it was considered that Tyr 8 was essential for GST catalysis. In contrast, we conclude that Tyr 8 facilitates the ionization of the thiol group of glutathione bound to glutathione S-transferase, but is not required for enzyme activity.  相似文献   

18.
Molecular Dynamics (MD) simulations were carried out for human acetylcholinesterase (hAChE) and its complex with Axillaridine–A, in order to dynamically explore the active site of the protein and the behaviour of the ligand at the peripheral binding site. Simulation of the enzyme alone showed that the active site of AChE is located at the bottom of a deep and narrow cavity whose surface is lined with rings of aromatic residues while Tyr72 is almost perpendicular to the Trp286, which is responsible for stable π -π interactions. The complexation of AChE with Axillaridine-A, results in the reduction of gorge size due to interaction between the ligand and the active site residues. The gorge size was determined by the distance between the center of mass of Glu81 and Trp286. As far as the geometry of the active site is concerned, the presence of ligand in the active site alters its specific conformation, as revealed by stable hydrogen bondings established between amino acids. With the increasing interaction between ligand and the active amino acids, size of the active site of the complex decreases with respect to time. Axillaridine-A, forms stable π -π interactions with the aromatic ring of Tyr124 that results in inhibition of catalytic activity of the enzyme. This π -π interaction keeps the substrate stable at the edge of the catalytic gorge by inhibiting its catalytic activity. The MD results clearly provide an explanation for the binding pattern of bulky steroidal alkaloids at the active site of AChE.  相似文献   

19.
Organophosphorus (OP) esters are known to bind covalently to the active site serine of enzymes in the serine hydrolase family. It was a surprise to find that proteins with no active site serine are also covalently modified by OP. The binding site in albumin, transferrin, and tubulin was identified as tyrosine. The goal of the present work was to determine whether binding to tyrosine is a general phenomenon. Fourteen proteins were treated with a biotin-tagged organophosphorus agent called FP-biotin. The proteins were digested with trypsin and the labeled peptides enriched by binding to monomeric avidin. Peptides were purified by HPLC and fragmented by collision induced dissociation in a tandem ion trap mass spectrometer. Eight proteins were labeled and six were not. Tyrosine was labeled in human alpha-2-glycoprotein 1 zinc-binding protein (Tyr 138, Tyr 174 and Tyr 181), human kinesin 3C motor domain (Tyr 145), human keratin 1 (Tyr 230), bovine actin (Tyr 55 and Tyr 200), murine ATP synthase beta (Tyr 431), murine adenine nucleotide translocase 1 (Tyr 81), bovine chymotrypsinogen (Tyr 201) and porcine pepsin (Tyr 310). Only 1–3 tyrosines per protein were modified, suggesting that the reactive tyrosine was activated by nearby residues that facilitated ionization of the hydroxyl group of tyrosine. These results suggest that OP binding to tyrosine is a general phenomenon. It is concluded that organophosphorus-reactive proteins include not only enzymes in the serine hydrolase family, but also proteins that have no active site serine. The recognition of a new OP-binding motif to tyrosine suggests new directions to search for mechanisms of long-term effects of OP exposure. Another application is in the search for biomarkers of organophosphorus agent exposure. Previous searches have been limited to serine hydrolases. Now proteins such as albumin and keratin can be considered.  相似文献   

20.
Heme oxygenase regiospecifically oxidizes heme at the alpha-meso position to give biliverdin IXalpha, CO, and iron. The heme orientation within the active site, which is thought to determine the oxidation regiospecificity, is shown here for the human enzyme (hHO1) to be largely determined by interactions between the heme carboxylic acid groups and residues Arg183 and Lys18 but not Tyr134. Mutation of either Arg183 or Lys18 individually does not significantly alter the NADPH-cytochrome P450 reductase-dependent reaction regiochemistry but partially shifts the oxidation to the beta/delta-meso positions in the reaction supported by ascorbic acid. Mutation of Glu29 to a lysine, which places a positive charge where it can interact with a heme carboxyl if the heme rotates by approximately 90 degrees, causes a slight loss of regiospecificity but combined with the R183E and K18E mutations results primarily in beta/delta-meso oxidation of the heme under all conditions. NMR analysis of heme binding to the triple K18E/E29K/R183E mutant confirms rotation of the heme in the active site. Kinetic studies demonstrate that mutations of Arg183 greatly impair the rate of the P450 reductase-dependent reaction, in accord with the earlier finding that Arg183 is involved in binding of the reductase to hHO1, but have little effect on the ascorbate reaction. Mutations of Asp140 and Tyr58 that disrupt the active site hydrogen bonding network impair catalytic rates but do not influence the oxidation regiochemistry. The results indicate both that the oxidation regiochemistry is largely controlled by ionic interactions of the heme propionic acid groups with the protein and that shifts in regiospecificity involve rotation of the heme about an axis perpendicular to the heme plane.  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