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 62 毫秒
1.
5,8-Linoleate diol synthase (5,8-LDS) of Aspergillus fumigatus was cloned, expressed, and compared with 7,8-LDS of the Take-all fungus. Replacements of Tyr and Cys in the conserved YRWH and FXXGPHXCLG sequences abolished 8R-dioxygenase (8-DOX) and hydroperoxide isomerase activities, respectively. The predicted α-helices of LDS were aligned with α-helices of cyclooxygenase-1 (COX-1) to identify the 8-DOX domains. N-terminal expression constructs of 5,8- and 7,8-LDS (674 of 1079, and 673 of 1165 residues), containing one additional α-helix compared to cyclooxygenase-1, yielded prominent 8R-DOX activities with apparently unchanged or slightly lower substrate affinities, respectively. Val-328 of 5,8-LDS did not influence the position of oxygenation in contrast to the homologous residues Val-349 of COX-1 and Leu-384 of 10R-dioxygenase. We conclude that ∼675 amino acids are sufficient to support 8-DOX activity.  相似文献   

2.
Aspergillus sp. contain ppoA, ppoB, and ppoC genes, which code for fatty acid oxygenases with homology to fungal linoleate 7,8-diol synthases (7,8-LDS) and cyclooxygenases. Our objective was to identify these enzymes, as ppo gene replacements show critical developmental aberrancies in sporulation and pathogenicity in the human pathogen Aspergillus fumigatus and the genetic model Aspergillus nidulans. The PpoAs of A. fumigatus and A. nidulans were identified as (8R)-dioxygenases with hydroperoxide isomerase activity, designated 5,8-LDS. 5,8-LDS transformed 18:2n-6 to (8R)-hydroperoxyoctadecadienoic acid ((8R)-HPODE) and (5S,8R)-dihydroxy-9Z,12Z-octadecadienoic acid ((5S,8R)-DiHODE). We also detected 8,11-LDS in A. fumigatus and (10R)-dioxygenases in both Aspergilli. The diol synthases oxidized [(8R)-(2)H]18:2n-6 to (8R)-HPODE with retention of the deuterium label, suggesting antarafacial hydrogen abstraction and insertion of molecular oxygen. Experiments with stereospecifically deuterated 18:2n-6 showed that (8R)-HPODE was isomerized by 5,8- and 8,11-LDS to (5S,8R)-DiHODE and to (8R,11S)-dihydroxy-9Z,12Z-octadecadienoic acid, respectively, by suprafacial hydrogen abstraction and oxygen insertion at C-5 and C-11. PpoCs were identified as (10R)-dioxygenases, which catalyzed abstraction of the pro-S hydrogen at C-8 of 18:2n-6, double bond migration, and antafacial insertion of molecular oxygen with formation of (10R)-hydroxy-8E,12Z-hydroperoxyoctadecadienoic acid ((10R)-HPODE). Deletion of ppoA led to prominent reduction of (8R)-H(P)ODE and complete loss of (5S,8R)-DiHODE biosynthesis, whereas biosynthesis of (10R)-HPODE was unaffected. Deletion of ppoC caused biosynthesis of traces of racemic 10-HODE but did not affect the biosynthesis of other oxylipins. We conclude that ppoA of Aspergillus sp. may code for 5,8-LDS with catalytic similarities to 7,8-LDS and ppoC for linoleate (10R)-dioxygenases. Identification of these oxygenases and their products will provide tools for analyzing the biological impact of oxylipin biosynthesis in Aspergilli.  相似文献   

3.
Linoleate diol synthases (LDS) are homologous 8(R)-dioxygenases with hydroperoxide isomerase activities, expressed in fungal pathogens of humanitarian importance. We report for the first time expression and site-directed mutagenesis of LDS. 7,8-LDS of the take-all fungus, expressed in Pichia pastoris, oxygenated 18:2n − 6 to 8(R)-hydroperoxylinoleic acid, which was unexpectedly isomerized to 5,8(R)-dihydroxylinoleic acid (60% 5S) and to 8(R),13-dihydroxyoctadeca-9(E),11(E)-dienoic acid. The latter was likely formed via hydrolysis of an unstable intermediate, 8(R),9(S)-epoxyoctadeca-10(E),12(Z)-dienoic acid. A tyrosyl radical is formed during 7,8-LDS catalysis, and Tyr376 is the sequence homolog to Tyr385 of cyclooxygenase-1. Tyr376Phe retained hydroperoxide isomerase activity but lacked 8(R)-dioxygenase activity. The putative proximal heme ligand His379 and the N-glycosylation site at Asn216 appeared to be critical for 8(R)-dioxygenase activity, as His379Gln and Asn216Gln were inactive. Treatment with α-mannosidase to shorten N- and O-linked mannosides inhibited the hydroperoxide isomerase but not the 8(R)-dioxygenase. Our results suggest that post-translational modifications may influence the oxidation mechanism of 7,8-LDS.  相似文献   

4.
HasA(SM) secreted by the Gram-negative bacterium Serratia marcescens belongs to the hemophore family. Its role is to take up heme from host heme carriers and to shuttle it to specific receptors. Heme is linked to the HasA(SM) protein by an unusual axial ligand pair: His32 and Tyr75. The nucleophilic nature of the tyrosine is enhanced by the hydrogen bonding of the tyrosinate to a neighboring histidine in the binding site: His83. We used isothermal titration microcalorimetry to examine the thermodynamics of heme binding to HasA(SM) and showed that binding is strongly exothermic and enthalpy driven: DeltaH = -105.4 kJ x mol(-1) and TDeltaS = -44.3 kJ x mol(-1). We used displacement experiments to determine the affinity constant of HasA(SM) for heme (K(a) = 5.3 x 10(10) M(-1)). This is the first time that this has been reported for a hemophore. We also analyzed the thermodynamics of the interaction between heme and a panel of single, double, and triple mutants of the two axial ligands His32 and Tyr75 and of His83 to assess the implication of each of these three residues in heme binding. We demonstrated that, in contrast to His32, His83 is essential for the binding of heme to HasA(SM), even though it is not directly coordinated to iron, and that the Tyr75/His83 pair plays a key role in the interaction.  相似文献   

5.
Zhang F  Gao J  Weng J  Tan C  Ruan K  Xu C  Jiang D 《Biochemistry》2005,44(2):719-725
To study its contribution to the assembly of the green plant manganese stabilizing protein (MSP) into photosystem II (PSII), tyrosine residues were specifically acetylated using N-acetylimidazole (NAI). In soluble MSP, three groups of Tyr residues could be differentiated by NAI acetylation: approximately 5 (actually approximately 5.2) Tyr residues could be easily acetylated (superficial), 1-2 Tyr residues could be acetylated when the NAI concentration was sufficiently high (superficially buried), and 1-2 Tyr residues could only be acetylated in the presence of the denaturant, urea (deeply buried). Acetylation of the 5.2 Tyr residues did not affect the reconstitution or oxygen-evolving activities of the MSP, and far-UV circular dichroism (CD) analysis showed that the altered MSP retained most of its native secondary structure. These results suggested that the 5.2 Tyr residues are not absolutely essential to the function of MSP. However, further modification of the 1-2 superficially buried Tyr residues (for a total acetylation of approximately 6.4 Tyr residues) completely abrogated the MSP rebinding and oxygen evolution activities. Finally, at least one tyrosine residue was inaccessible to NAI until MSP was completely unfolded by 8 M urea. Deacetylation of MSP with 6.4 or 8 acetylated Tyr residues with hydroxylamine restored most of the rebinding and oxygen-evolving activities. A prominent red shift in fluorescence spectra of MSP (excited at 280 or 295 nm) was observed after modification of 6.4 Tyr residues, and a further shift could be found after all 8 Tyr residues were modified, indicating a great loss of native secondary structure. Far-UV CD revealed that MSP was mostly unfolded when 6.4 Tyr residues were modified and completely unfolded when all 8 Tyr residues were modified. Fluorescence and far-UV CD studies revealed that loss of MSP rebinding to PSII membranes following NAI modification correlated well with conformational changes in MSP. Together, these results indicate that different tyrosine residues have different contributions to the binding and assembly of MSP into PSII.  相似文献   

6.
The structure of aldehyde reductase (ALR1) in ternary complex with the coenzyme NADPH and 3,5-dichlorosalicylic acid (DCL), a potent inhibitor of human 20α-hydroxysteroid dehydrogenase (AKR1C1), was determined at a resolution of 2.41 Å. The inhibitor formed a network of hydrogen bonds with the active site residues Trp22, Tyr50, His113, Trp114 and Arg312. Molecular modelling calculations together with inhibitory activity measurements indicated that DCL was a less potent inhibitor of ALR1 (256-fold) when compared to AKR1C1. In AKR1C1, the inhibitor formed a 10-fold stronger binding interaction with the catalytic residue (Tyr55), non-conserved hydrogen bonding interaction with His222, and additional van der Waals contacts with the non-conserved C-terminal residues Leu306, Leu308 and Phe311 that contribute to the inhibitor’s selectivity advantage for AKR1C1 over ALR1.  相似文献   

7.
Using a combination of DNA and hybrid DNA-RNA substrates, we have analyzed the mechanism of phosphoryl transfer by the Flp site-specific recombinase in three different reactions: DNA strand breakage and joining, and two types of RNA cleavage activities. These reactions were then used to characterize Flp variants altered at His309 and His345, amino acid residues that are in close proximity to two key catalytic residues (Arg308 and Tyr343). These histidine residues are important for strand cutting by Tyr343, the active-site nucleophile of Flp, but neither residue contributes to the type II RNA cleavage activity or to the strand-joining reaction in a pre-cleaved substrate. Strand cleavage reactions using small, diffusible nucleophiles indicate that this histidine pair contributes to the correct positioning and activation of Tyr343 within the shared active site of Flp. The implications of these results are evaluated against the recently solved crystal structure of Flp in association with a Holliday junction.  相似文献   

8.
P Canioni  P J Cozzone 《Biochimie》1979,61(3):343-354
The low-field portion of the 360 MHz proton N.M.R. spectrum of native porcine pancreatic colipase has been studied as a function of pH over the pH range 2-12. Resonances associated with the 26 protons of the aromatic rings of the two histidines, two phenylalanines and three tyrosines have been identified and tentatively assigned to specific residues. Titrations of pH yielded apparent pKa's of 7.9, 6.9, 10.4, 10.3 and 11.3 for His I (His 30), His II (His 86), Tyr I (Tyr 56 or 57), Tyr II (Tyr 56 or 57) and Tyr III (Tyr 53) respectively (tentative assignments). The high pKa value of His 30 is attributed to the vicinity of Asp 31. The mobility of the aromatic ring of Tyr 53 is hindered and an upper bound of 500 s-1 on the rate of rotation can be estimated. The aromatic rings of the 2 other tyrosine residues and of the 2 phenylalanine residues can rotate freely on the N.M.R. time scale. The study of perturbations in titration profiles and chemical shift values reveals a specific interaction of His 86 with Tyr I and, to a lesser extent, Tyr II. The existence of this interaction indicates that the protein folding brings in close spatial vicinity two distant regions of the covalent structure to form a "hydrophobic-aromatic" site which might be involved in the binding of bile salt micelles to pancreatic colipase.  相似文献   

9.
Linoleate (10R)-dioxygenase (10R-DOX) of Aspergillus fumigatus was cloned and expressed in insect cells. Recombinant 10R-DOX oxidized 18:2n-6 to (10R)-hydroperoxy-8(E),12(Z)-octadecadienoic acid (10R-HPODE; ∼90%), (8R)-hydroperoxylinoleic acid (8R-HPODE; ∼10%), and small amounts of 12S(13R)-epoxy-(10R)-hydroxy-(8E)-octadecenoic acid. We investigated the oxygenation of 18:2n-6 at C-10 and C-8 by site-directed mutagenesis of 10R-DOX and 7,8-linoleate diol synthase (7,8-LDS), which forms ∼98% 8R-HPODE and ∼2% 10R-HPODE. The 10R-DOX and 7,8-LDS sequences differ in homologous positions of the presumed dioxygenation sites (Leu-384/Val-330 and Val-388/Leu-334, respectively) and at the distal site of the heme (Leu-306/Val-256). Leu-384/Val-330 influenced oxygenation, as L384V and L384A of 10R-DOX elevated the biosynthesis of 8-HPODE to 22 and 54%, respectively, as measured by liquid chromatography-tandem mass spectrometry analysis. The stereospecificity was also decreased, as L384A formed the R and S isomers of 10-HPODE and 8-HPODE in a 3:2 ratio. Residues in this position also influenced oxygenation by 7,8-LDS, as its V330L mutant augmented the formation of 10R-HPODE 3-fold. Replacement of Val-388 in 10R-DOX with leucine and phenylalanine increased the formation of 8R-HPODE to 16 and 36%, respectively, whereas L334V of 7,8-LDS was inactive. Mutation of Leu-306 with valine or alanine had little influence on the epoxyalcohol synthase activity. Our results suggest that Leu-384 and Val-388 of 10R-DOX control oxygenation of 18:2n-6 at C-10 and C-8, respectively. The two homologous positions of prostaglandin H synthase-1, Val-349 and Ser-353, are also critical for the position and stereospecificity of the cyclooxygenase reaction.Linoleate diol synthases (LDS)2 and linoleate 10R-DOX are fungal fatty acid dioxygenases of the myeloperoxidase gene family (1-3). LDS have dual enzyme activities and transform 18:2n-6 sequentially to 8R-HPODE in an 8R-dioxygenase reaction and to 5,8-, 7,8-, or 8,11-DiHODE in hydroperoxide isomerase reactions. These oxylipins affect sporulation, development, and pathogenicity of Aspergilli (4-6). Fatty acid dioxygenases of the myeloperoxidase gene family also occur in vertebrates, plants, and algae (7-9). The most thoroughly investigated vertebrate enzymes are ovine PGHS-1 and mouse PGHS-2 with known crystal structures (10-12). PGHS transforms 20:4n-6 to PGG2 in a cyclooxygenase and PGG2 to PGH2 in a peroxidase reaction. Aspirin and other nonsteroidal anti-inflammatory drugs inhibit the cyclooxygenase reaction. This is of paramount medical importance (13, 14), and PGHS-1 and -2 are commonly known as COX-1 and -2 (15). α-DOX occur in plants and algae, and biosynthesis of α-DOX in plants is elicited by pathogens (7). α-DOX oxidizes fatty acids to unstable (2R)-hydroperoxides, which readily break down nonenzymatically to fatty acid aldehydes and CO2 (7).LDS, 10R-DOX, PGHS, and α-DOX oxygenate fatty acids to different products, but their oxygenation mechanisms have mechanistic similarities. Sequence alignment shows that many critical amino acid residues for the cyclooxygenase reaction are conserved in LDS, 10R-DOX, and α-DOX. These include the proximal histidine heme ligand, the distal histidine, and the catalytic important tyrosine (Tyr-385) of PGHS-1. The latter is oxidized to a tyrosyl radical, which initiates the cyclooxygenase reaction by abstraction of the pro-S hydrogen at C-13 of 20:4n-6 (16). In analogy, LDS and 10R-DOX catalyze stereospecific abstraction of the pro-S hydrogen at C-8 of 18:2n-6 (3), whereas α-DOX abstracts the pro-R hydrogen at C-2 of fatty acids (17). Site-directed mutagenesis of the conserved tyrosine homologues of Tyr-385 and proximal heme ligands abolishes the dioxygenase activities of 7,8-LDS and α-DOX (17, 18). The orientation of the substrate at the dioxygenation site differs. The carboxyl groups of fatty acids are positioned in a hydrophobic grove close to the tyrosine residue of α-DOX (19). In contrast, the ω ends of eicosanoic fatty acids are buried deep inside the cyclooxygenase channel so that C-13 lies in the vicinity of Tyr-385 (20). Several observations suggest that 18:2n-6 may also be positioned with its ω end embedded in the interior of 7,8-LDS of Gaeumannomyces graminis (18).7,8-LDS of G. graminis and Magnaporthe grisea and 5,8-LDS of Aspergillus nidulans have been sequenced (5, 8, 21). Gene targeting revealed the catalytic properties of 5,8-LDS, 8,11-LDS, and 10R-DOX in Aspergillus fumigatus and A. nidulans (3). Homologous genes can be found in other Aspergilli spp. Alignment of the two 7,8-LDS amino acid sequences with 5,8-LDS, 8,11-LDS, and 10R-DOX sequences of five Aspergilli revealed several conserved regions with single amino acid differences between the enzymes with 8R-DOX and 10R-DOX activities, as illustrated by the selected sequences in Fig. 1. Leu-306, Leu-384, and Val-388 of 10R-DOX are replaced in 5,8- and 7,8-LDS by valine, valine, and leucine residues, respectively. Whether these amino acids are important for the oxygenation mechanism is unknown, and this is one topic of the present investigation. The predicted secondary structure of 10R-DOX suggests that Leu-384 of 10R-DOX can be present in an α-helix with Val-388 close to its border. This α-helix is homologous to helix 6 of PGHS-1, which contains Val-349 and Ser-353 at the homologous positions of Leu-384 and Val-388 (Fig. 1).Open in a separate windowFIGURE 1.Alignments of partial amino acid sequences of five heme containing fatty acid dioxgenases and a comparison of the predicted secondary structure of 10R-DOX with ovine PGHS-1. A, top, amino acids residues at the presumed peroxidase and hydroperoxide isomerase sites. The last two residues, His and Asn, are conserved in all myeloperoxidases (1). Middle and bottom, amino acid residues of the presumed dioxygenation sites are shown. Conserved residues in all sequences are in boldface, and mutated residues of 10R-DOX and/or 7,8-LDS are marked by an asterisk. B, alignment of partial amino acid sequences of 10R-DOX with ovine PGHS-1, and a secondary structure prediction of the 10R-DOX sequence. The secondary structure of 10R-DOX was predicted by PSIPRED (43) and the secondary structure of ovine PGHS-1 from its crystal structure (Protein Data Bank code 1diy; cf. Ref 19). In short, our first strategy for site-directed mutagenesis was to switch hydrophobic residues between the enzymes with 10R- and 8R-DOX activities and to assess the effects on the DOX and hydroperoxide isomerase activities (10R-DOX/7,8-LDS: Leu-306/Val-256, Leu-384/Val-330, Val-388/Leu-334, and Ala-426/Ile-375) and to switch one hydrophobic/charged residue (Ala-435/Glu-384). Only catalytically active pairs would provide clear information on their importance for the position of dioxygenation (e.g. L384V of 10R-DOX and V330L of 7,8-LDS, both of which were active). Unfortunately, replacements of 7,8-LDS often led to inactivation or very low activity (e.g. V330A, V330M, I375A, E384A). Our second strategy was to study replacements in two homologous positions of ovine PGHS-1 (Val-349 and Ser-353) with smaller and larger hydrophobic residues, i.e. at Leu-384 and Val-388 of 10R-DOX. Abbreviations used are as follows: oCOX-1, ovine cyclooxygenase-1; Af, A. fumigatus; Gg, G. graminis. The GenBank™ protein sequences were derived from P05979, EAL89712, AAD49559, EAL84400, and ACL14177. The amino acid sequences were aligned with the ClustalW algorithm (DNAStar).The overall three-dimensional structures of myeloperoxidases are conserved. It is therefore conceivable that important residues for substrate binding in the cyclooxygenase channel of PGHS could be conserved in LDS and 10R-DOX. The three-dimensional structure of ovine PGHS-1 shows that Val-349 and Ser-353 are close to C-3 and C-4 of 20:4n-6, and residues in these positions can alter both position and stereospecificity of oxygenation (22-24). Replacement of Val-349 of PGHS-1 with alanine increased the biosynthesis of 11R-HETE, whereas V349L decreased the generation of 11R-H(P)ETE and increased formation of 15(R/S)-H(P)ETE (23, 25). V349I formed PGG2 with 15R configuration (22, 24). Replacement of Ser-353 with threonine reduced cyclooxygenase and peroxidase activities by over 50% and increased the biosynthesis of 11R-HPETE and 15S-HPETE 4-5 times (23).There is little information on the hydroperoxide isomerase and peroxidase sites of LDS (18, 26), but the latter could be structurally related to the peroxidase site of PGHS. PGG2 and presumably 8R-HPODE bind to the distal side of the heme group, which can be delineated by hydrophobic amino acid residues (27). Val-291 is one of these residues, which form a dome over the distal heme side of COX-1. The V291A mutant retained cyclooxygenase and peroxidase activities (27). 5,8- and 7,8-LDS also have valine residues in the homologous position, whereas 8,11-LDS and 10R-DOX have leucine residues (Fig. 1). Whether these hydrophobic residues are important for the peroxidase activities is unknown.In this study we decided to compare the two catalytic sites of 10R-DOX of A. fumigatus and 7,8-LDS (EC 1.13.11.44) of G. graminis (18). Our first aim was to find a robust expression system for 10R-DOX of A. fumigatus. The second objective was to determine whether C16 and C20 fatty acid substrates enter the oxygenation site of 10R-DOX “head” or “tail” first. Unexpectedly, we found that 10R-DOX oxygenated 20:4n-6 by hydrogen abstraction at both C-13 and C-10 with formation of two nonconjugated and four cis-trans-conjugated HPETEs. Our third objective was to investigate the structural differences between 10R-DOX and 7,8-LDS of G. graminis, which could explain that oxygenation of 18:2n-6 mainly occurred at C-10 and at C-8, respectively. The strategy for site-directed mutagenesis of 10R-DOX and 7,8-LDS is outlined in the legend to Fig. 1; an alignment of the amino acid sequences of 10R-DOX and 7,8-LDS is found in supplemental material.  相似文献   

10.
Structural aspects of the binding of the linear ligands N alpha-acetyl-L-lysine (AcLys) and epsilon-aminocaproic acid (epsilon ACA) and of the cyclic analogs trans-(aminomethyl)-cyclohexanecarboxylic acid (AMCHA) and p-benzylaminesulfonic acid (BASA) to the intact plasminogen kringle 4 domain have been investigated by 1H-NMR spectroscopy at 300 and 600 MHz. Ligand binding results in consistent shifts of the His-II (His31), Trp-I (Trp25?), Trp-II (Trp62?), Trp-III (Trp72), Tyr-II (Tyr50), and Phe64 ring signals. BASA tends to induce larger shifts than elicited by the aliphatic ligands, most noticeably on Trp-II and on Trp72, suggesting that the ligand aromatic ring interacts with the two indole groups. Trp-II and, to lesser extent, Trp-I interact with an acidic side chain group, in a manner that is blocked by BASA. BASA binding also perturbs Tyr-II (Tyr50), Tyr-III (Tyr41), and Tyr-IV (Tyr74) over a wide pH range and lowers the pKa* of His31 from approximately 4.8 to approximately 4.6. His-III (His33) responds to BASA and AMCHA but is relatively insensitive to the linear ligands. His33 carries a sterically shielded side chain which, in conjunction with Leu46, Trp-I, Tyr50, and Tyr74, participates in structuring the kringle hydrophobic core, contiguous to the binding site. Pronounced shifts are observed for aliphatic resonances stemming from the kringle-bound molecules of AMCHA, AcLys, and epsilon ACA. It is proposed that the lysine-binding site is mostly supported by the loop that extends from Cys51 through Cys71 and that aromatic residues, which include Trp-II, Trp72, and Phe64, play a major role in interacting with the nonpolar segment of the ligand molecule. The binding site also encompasses Tyr50, Tyr74, His31, and His33 although it is not clear the extent to which these residues interact directly with the ligand.  相似文献   

11.
Among manganese superoxide dismutases, residues His30 and Tyr174 are highly conserved, forming part of the substrate access funnel in the active site. These two residues are structurally linked by a strong hydrogen bond between His30 NE2 from one subunit and Tyr174 OH from the other subunit of the dimer, forming an important element that bridges the dimer interface. Mutation of either His30 or Tyr174 in Escherichia coli MnSOD reduces the superoxide dismutase activity to 30--40% of that of the wt enzyme, which is surprising, since Y174 is quite remote from the active site metal center. The 2.2 A resolution X-ray structure of H30A-MnSOD shows that removing the Tyr174-->His30 hydrogen bond from the acceptor side results in a significant displacement of the main-chain segment containing the Y174 residue, with local rearrangement of the protein. The 1.35 A resolution structure of Y174F-MnSOD shows that disruption of the same hydrogen bond from the donor side has much greater consequences, with reorientation of F174 having a domino effect on the neighboring residues, resulting in a major rearrangement of the dimer interface and flipping of the His30 ring. Spectroscopic studies on H30A, H30N, and Y174F mutants show that (like the previously characterized Y34F mutant of E. coli MnSOD) all lack the high pH transition of the wt enzyme. This observation supports assignment of the pH sensitivity of MnSOD to coordination of hydroxide ion at high pH rather than to ionization of the phenolic group of Y34. Thus, mutations near the active site, as in the Y34F mutant, as well as at remote positions, as in Y174F, similarly affect the metal reactivity and alter the effective pK(a) for hydroxide ion binding. These results imply that hydrogen bonding of the H30 imidazole N--H group plays a key role in substrate binding and catalysis.  相似文献   

12.
The 3D structure of a novel epoxide hydrolase from Aspergillus niger SQ-6 (sqEH) was constructed by using homology modeling and molecular dynamics simulations. Based on the 3D model, Asp191, His369 and Glu343 were predicted as catalytic triad. The putative active pocket is a hydrophobic environment and is rich in some important non—polar residues (Pro318, Trp282, Pro319, Pro317 and Phe242). Using three sets of epoxide inhibitors for docking study, the interaction energies of sqEH with each inhibitor are consistent with their inhibitory effects in previous experiments. Moreover, a critical water molecule which closes to the His369 was identified to be an ideal position for the hydrolysis step of the reaction. Two tyrosine residues (Tyr249 and Tyr312) are able to form hydrogen bonds with the epoxide oxygen atom to maintain the initial binding and positioning of the substrate in the active pocket. These docked complex models can well interpret the substrate specificity of sqEH, which could be relevant for the structural—based design of specific epoxide inhibitors. Figure    相似文献   

13.
Among the aromatic residues in protein structures, histidine (His) is unique, as it can exist in the neutral or positively charged form at the physiological pH. As such, it can interact with other aromatic residues as well as form hydrogen bonds with polar and charged (both negative and positive) residues. We have analyzed the geometry of interaction of His residues with nine other planar side chains containing aromatic (residues Phe, Tyr, Trp, and His), carboxylate (Asp and Glu), carboxamide (Asn and Gln) and guanidinium (Arg) groups in 432 polypeptide chains. With the exception of the aspartic (Asp) and glutamic (Glu) acid side-chains, all other residues prefer to interact in a face-to-face or offset-face-stacked orientation with the His ring. Such a geometry is different from the edge-to-face relative orientation normally associated with the aromatic-aromatic interaction. His-His pair prefers to interact in a face-to-face orientation; however, when both the residues bind the same metal ion, the interplanar angle is close to 90 degrees. The occurrence of different interactions (including the nonconventional N-H...pi and C-H...pi hydrogen bonds) have been correlated with the relative orientations between the interacting residues. Several structural motifs, mostly involved in binding metal ions, have been identified by considering the cases where His residues are in contact with four other planar moieties. About 10% of His residues used here are also found in sequence patterns in PROSITE database. There are examples of the amino end of the Lys side chain interacting with His residues in such a way that it is located on an arc around a ring nitrogen atom.  相似文献   

14.
GPR40 was formerly an orphan G protein-coupled receptor whose endogenous ligands have recently been identified as free fatty acids (FFAs). The receptor, now named FFA receptor 1, has been implicated in the pathophysiology of type 2 diabetes and is a drug target because of its role in FFA-mediated enhancement of glucose-stimulated insulin release. Guided by molecular modeling, we investigated the molecular determinants contributing to binding of linoleic acid, a C18 polyunsaturated FFA, and GW9508, a synthetic small molecule agonist. Twelve residues within the putative GPR40-binding pocket including hydrophilic/positively charged, aromatic, and hydrophobic residues were identified and were subjected to site-directed mutagenesis. Our results suggest that linoleic acid and GW9508 are anchored on their carboxylate groups by Arg(183), Asn(244), and Arg(258). Moreover, His(86), Tyr(91), and His(137) may contribute to aromatic and/or hydrophobic interactions with GW9508 that are not present, or relatively weak, with linoleic acid. The anchor residues, as well as the residues Tyr(12), Tyr(91), His(137), and Leu(186), appear to be important for receptor activation also. Interestingly, His(137) and particularly His(86) may interact with GW9508 in a manner dependent on its protonation status. The greater number of putative interactions between GPR40 and GW9508 compared with linoleic acid may explain the higher potency of GW9508.  相似文献   

15.
Generation of hydrogen peroxide and hydroxyl radicals in L-amino acid solutions in phosphate buffer, pH 7.4, under X-ray irradiation was determined by enhanced chemiluminescence in the luminol-p-iodophenol-peroxidase system and using the fluorescent probe coumarin-3-carboxylic acid, respectively. Amino acids are divided into three groups according to their effect on the hydrogen peroxide formation under irradiation: those decreasing yield of H2O2, having no effect, and increasing its yield. All studied amino acids at 1 mM concentration decrease the yield of hydroxyl radicals in solution under X-ray irradiation. However, the highest effect is observed in the order: Cys > His > Phe = Met = Trp > Tyr. At Cys, Tyr, and His concentrations close to physiological, the yield of hydroxyl radicals decreases significantly. Immunoenzyme analysis using monoclonal antibodies to 8-oxoguanine (8-oxo-7,8-dihydroguanine) was applied to study the effect of amino acids with the most pronounced antioxidant properties (Cys, Met, Tyr, Trp, Phe, His, Lys, Arg, Pro) on 8-oxoguanine formation in vitro under X-ray irradiation. It is shown that amino acids decrease the content of 8-oxoguanine in DNA. These amino acids within DNA-binding proteins may protect intracellular DNA against oxidative damage caused by formation of reactive oxygen species in conditions of moderate oxidative stress.  相似文献   

16.
Haruta N  Aki M  Ozaki S  Watanabe Y  Kitagawa T 《Biochemistry》2001,40(23):6956-6963
Conformational change of myoglobin (Mb) accompanied by binding of a ligand was investigated with 244 nm excited ultraviolet resonance Raman Spectroscopy (UVRR). The UVRR spectra of native sperm whale (sw) and horse (h) Mbs and W7F and W14F swMb mutants for the deoxy and CO-bound states enabled us to reveal the UVRR spectra of Trp7, Trp14, and Tyr151 residues, separately. The difference spectra between the deoxy and CO-bound states reflected the environmental or structural changes of Trp and Tyr residues upon CO binding. The W3 band of Trp7 near the N-terminus exhibited a change upon CO binding, while Trp14 did not. Tyr151 in the C-terminus also exhibited a definite change upon CO binding, but Tyr103 and Tyr146 did not. The spectral change of Tyr residues was characterized through solvent effects of a model compound. The corresponding spectral differences between CO- and n-butyl isocyanide-bound forms were much smaller than those between the deoxy and CO-bound forms, suggesting that the conformation change in the C- and N-terminal regions is induced by the proximal side of the heme through the movement of iron. Although the swinging up of His64 upon binding of a bulky ligand is noted by X-ray crystallographic analysis, UVRR spectra of His for the n-butyl isocyanide-bound form did not detect the exposure of His64 to solvent.  相似文献   

17.
The crystal structure of a chitinase from Carica papaya has been solved by the molecular replacement method and is reported to a resolution of 1.5 A. This enzyme belongs to family 19 of the glycosyl hydrolases. Crystals have been obtained in the presence of N-acetyl- d-glucosamine (GlcNAc) in the crystallization solution and two well-defined GlcNAc molecules have been identified in the catalytic cleft of the enzyme, at subsites -2 and +1. These GlcNAc moieties bind to the protein via an extensive network of interactions which also involves many hydrogen bonds mediated by water molecules, underlying their role in the catalytic mechanism. A complex of the enzyme with a tetra-GlcNAc molecule has been elaborated, using the experimental interactions observed for the bound GlcNAc saccharides. This model allows to define four major substrate interacting regions in the enzyme, comprising residues located around the catalytic Glu67 (His66 and Thr69), the short segment E89-R90 containing the second catalytic residue Glu89, the region 120-124 (residues Ser120, Trp121, Tyr123, and Asn124), and the alpha-helical segment 198-202 (residues Ile198, Asn199, Gly201, and Leu202). Water molecules from the crystal structure were introduced during the modeling procedure, allowing to pinpoint several additional residues involved in ligand binding that were not previously reported in studies of poly-GlcNAc/family 19 chitinase complexes. This work underlines the role played by water-mediated hydrogen bonding in substrate binding as well as in the catalytic mechanism of the GH family 19 chitinases. Finally, a new sequence motif for family 19 chitinases has been identified between residues Tyr111 and Tyr125.  相似文献   

18.
An NMR study has been made of porcine pancreatic phospholipase A2 (PLA) in three environments: free in solution, in a binary complex with dodecylphosphocholine micelles, and in a ternary complex with a micelle and the substrate-like inhibitor (R)-1-octyl-2-(N-dodecanoylamino)-2-deoxyglycero-3-phosph oglycol. 1H and 15N chemical shifts, amide exchange rates, and NOE intensities are compared for the enzyme in different environments. From these data, structural differences are found for the N-terminal part, the end of the surface loop at residues Tyr69 and Thr70, and the active site residue His48, and also for the Ca-binding loop (residues 28-32). Specifically, when binding to a micelle, the side chains of residues Ala1, Trp3, and Tyr69, as well as all protons of Thr70, are found to be closer together. After subsequent introduction of the competitive inhibitor, further changes are found for these residues. The N-terminus is flexible in PLA free in solution, in contrast with the crystal structures where it adopts an alpha-helical conformation. According to the NMR data, this helix is rigidly formed only in the ternary complex. Furthermore, in the ternary complex, the N-terminal amino group and the exchangeable hydrogen at N3 of the ring of His48 are observed. We propose that PLA is activated in two steps. An initial conformational change occurs upon binding to a micellar interface. The catalytically active conformation of the enzyme, which has an extensive network of hydrogen bonds, is formed only when binding a substrate or competitive inhibitor at a lipid-water interface.  相似文献   

19.
A quadruple mutant of sperm whale myoglobin was constructed to mimic the structure found in Ascaris suum hemoglobin. The replacements include His(E7)-->Gln, Leu(B10)-->Tyr, Thr(E10)--> Arg, and Ile(G8)-->Phe. Single, double, and triple mutants were characterized to dissect out the effects of the individual substitutions. The crystal structures of the deoxy and oxy forms of the quadruple mutant were determined and compared with that of native Ascaris hemoglobin. Tyr(B10) myoglobin displays low O(2) affinity, high dissociation rate constants, and heterogeneous kinetic behavior, suggesting unfavorable steric interactions between the B10 phenol side chain and His(E7). In contrast, all mutants containing the Tyr(B10)/Gln(E7) pair show high O(2) affinity, low dissociation rate constants, and simple, monophasic kinetic behavior. Replacement of Ile(107) with Phe enhances nanosecond geminate recombination singly and in combination with the Tyr(B10)/Gln(E7)/Arg(E10) mutation by limiting access to the Xe4 site. These kinetic results and comparisons with native Ascaris hemoglobin demonstrate the importance of distal pocket cavities in governing the kinetics of ligand binding. The approximately 150-fold higher O(2) affinity of Ascaris hemoglobin compared with that for Tyr(B10)/Gln(E7)-containing myoglobin mutants appears to be the result of favorable proximal effects in the Ascaris protein, due to a staggered orientation of His(F8), the lack of a hydrogen bonding lattice between the F4, F7, and F8 residues, and the presence of a large polar Trp(G5) residue in the interior portion of the proximal heme pocket.  相似文献   

20.
Chemical modification of ovomucoid third domain (DIII) has been conducted to characterize the binding site residues that determine antigenecity and allergenecity of DIII. Nitration of Tyr, ethoxyformylation of His and succinylation of Lys residues led to a decrease of alpha-helix content of DIII. Modification of His, Tyr, Glu, Asp and Lys residues on DIII resulted in a reduction of human IgG binding activity, but little effect on IgE binding activity. These results suggest that hydrophilic residues appear to be more critical for human IgG binding site, whereas hydrophobic residues may be more important for IgG binding site.  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