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.
C-C hydrolase MhpC (2-hydroxy-6-keto-nona-1,9-dioic acid 5,6-hydrolase) from Escherichia coli catalyses the hydrolytic C-C cleavage of the meta-ring fission product on the phenylpropionic acid catabolic pathway. The crystal structure of E. coli MhpC has revealed a number of active-site amino acid residues that may participate in catalysis. Site-directed mutants of His263, Ser110, His114, and Ser40 have been analysed using steady-state and stopped-flow kinetics. Mutants H263A, S110A and S110G show 10(4)-fold reduced catalytic efficiency, but still retain catalytic activity for C-C cleavage. Two distinct steps are observed by stopped-flow UV/Vis spectrophotometry, corresponding to ketonisation and C-C cleavage: H263A exhibits very slow ketonisation and C-C cleavage, whereas S110A and S110G exhibit fast ketonisation, an intermediate phase, and slow C-C cleavage. H114A shows only twofold-reduced catalytic efficiency, ruling out a catalytic role, but shows a fivefold-reduced K(M) for the natural substrate, and an ability to process an aryl-containing substrate, implying a role for His114 in positioning of the substrate. S40A shows only twofold-reduced catalytic efficiency, but shows a very fast (500 s(-1)) interconversion of dienol (317 nm) to dienolate (394 nm) forms of the substrate, indicating that the enzyme accepts the dienol form of the substrate. These data imply that His263 is responsible for both ketonisation of the substrate and for deprotonation of water for C-C cleavage, a novel catalytic role in a serine hydrolase. Ser110 has an important but non-essential role in catalysis, which appears not to be to act as a nucleophile. A catalytic mechanism is proposed involving stabilisation of reactive intermediates and activation of a nucleophilic water molecule by Ser110.  相似文献   

2.
3.
2-Hydroxy-6-ketonona-2,4-diene-1,9-dioic acid 5,6-hydrolase (MhpC) is a 62 kDa homodimeric enzyme of the phenylpropionate degradation pathway of Escherichia coli. The 2.1 A resolution X-ray structure of the native enzyme determined from orthorhombic crystals confirms that it is a member of the alpha/beta hydrolase fold family, comprising eight beta-strands interconnected by loops and helices. The 2.8 A resolution structure of the enzyme co-crystallised with the non-hydrolysable substrate analogue 2,6-diketo-nona-1,9-dioic acid (DKNDA) confirms the location of the active site in a buried channel including Ser110, His263 and Asp235, postulated contributors to a serine protease-like catalytic triad in homologous enzymes. It appears that the ligand binds in two separate orientations. In the first, the C6 keto group of the inhibitor forms a hemi-ketal adduct with the Ser110 side-chain, the C9 carboxylate group interacts, via the intermediacy of a water molecule, with Arg188 at one end of the active site, while the C1 carboxylate group of the inhibitor comes close to His114 at the other end. In the second orientation, the C1 carboxylate group binds at the Arg188 end of the active site and the C9 carboxylate group at the His114 end. These arrangements implicated His114 or His263 as plausible contributors to catalysis of the initial enol/keto tautomerisation of the substrate but lack of conservation of His114 amongst related enzymes and mutagenesis results suggest that His263 is the residue involved. Variability in the quality of the electron density for the inhibitor amongst the eight molecules of the crystal asymmetric unit appears to correlate with alternative positions for the side-chain of His114. This might arise from half-site occupation of the dimeric enzyme and reflect the apparent dissociation of approximately 50% of the keto intermediate from the enzyme during the catalytic cycle.  相似文献   

4.
Oxidation of cis-3,4-dehydroadipyl-CoA semialdehyde to cis-3,4-dehydroadipyl-CoA by the aldehyde dehydrogenase, ALDH(C) (EC.1.2.1.77), is an essential step in the metabolism of benzoate in Burkholderia xenovorans LB400. In a previous study, we established a structural blueprint for this novel group of ALDH enzymes. Here, we build significantly on this initial work and propose a detailed reaction mechanism for ALDH(C) based on comprehensive structural and functional investigations of active site residues. Kinetic analyses reveal essential roles for C296 as the nucleophile and E257 as the associated general base. Structural analyses of E257Q and C296A variants suggest a dynamic charge repulsion relationship between E257 and C296 that contributes to the inherent flexibility of E257 in the native enzyme, which is further regulated by E496 and E167. A proton relay network anchored by E496 and supported by E167 and K168 serves to reset E257 for the second catalytic step. We also propose that E167, which is unique to ALDH(C) and its homologs, serves a critical role in presenting the catalytic water to the newly reset E257 such that the enzyme can proceed with deacylation and product release. Collectively, the reaction mechanism proposed for ALDH(C) promotes a greater understanding of these novel ALDH enzymes, the ALDH super-family in general, and benzoate degradation in B. xenovorans LB400.  相似文献   

5.
Li JJ  Li C  Blindauer CA  Bugg TD 《Biochemistry》2006,45(41):12461-12469
C-C hydrolase enzymes MhpC and BphD catalyze the hydrolytic C-C cleavage of meta-ring fission intermediates on the Escherichia coli phenylpropionic acid and Burkholderia xenovorans LB400 biphenyl degradation pathways and are both members of the alpha/beta-hydrolase family containing a Ser-His-Asp catalytic triad. The catalytic mechanism of this family of enzymes is thought to proceed via a gem-diol reaction intermediate, which has not been observed directly. Site-directed single mutants of BphD in which catalytic residues His-265 and Ser-112 were replaced with Ala were found to possess 10(4)-fold reduced k(cat) values, and in each case, the C-C cleavage step was shown by pre-steady-state kinetic analysis to be rate-limiting. The processing of a 6-(13)C-labeled aryl-containing substrate by these H265A or S112A mutant BphD enzymes was monitored directly by (13)C NMR spectroscopy. A new line-broadened signal was observed at 128 ppm for each enzyme, corresponding to the proposed gem-diol reaction intermediate, over a time scale of 1-24 h. A similar signal was observed upon incubation of the (13)C-labeled substrate with an H114A MhpC mutant, which is able to accept the 6-phenyl-containing substrate, on a shorter time scale. The direct observation of a gem-diol intermediate provides further evidence that supports a general base mechanism for this family of enzymes.  相似文献   

6.
7.
8.
Goodall JJ  Chen GJ  Page MG 《Biochemistry》2004,43(15):4583-4591
The peptidyl-tRNA hydrolase (Pth) enzyme plays an essential role in recycling tRNA from peptidyl-tRNA that has prematurely dissociated from the ribosome. In this study of Escherichia coli Pth, the critical role of histidine 20 was investigated by site-directed mutagenesis, stopped-flow kinetic measurements, and chemical modification. The histidine residue at position 20 is known to play an important role in the hydrolysis reaction, but stopped-flow fluorescence measurements showed that, although the His20Asn Pth mutant enzyme was unable to hydrolyze the substrate, the enzyme retained the ability to bind peptidyl-tRNA. Chemical modification of Pth with diethyl pyrocarbonate (DEPC) showed that a residue, with a pK(a) value of 6.3, was essential for substrate hydrolysis and that the stoichiometry of inhibition was 0.70 +/- 0.06 mol of DEPC/mol of enzyme, indicating that modification of only a single residue by DEPC was responsible for the loss of activity. Parallel chemical modification studies with the His20Asn and Asp93Asn mutant enzymes showed that this essential residue was His20. These studies indicate that histidine 20 acts as the catalytic base in the hydrolysis of peptidyl-tRNA by Pth.  相似文献   

9.
Previous work has shown that the C-terminal portion of BphA, especially two amino acid segments designated region III and region IV, influence the regiospecificity of the biphenyl dioxygenase (BPDO) toward 2,2'-dichlorobiphenyl (2,2'-CB). In this work, we evolved BPDO by shuffling bphA genes amplified from polychlorinated biphenyl-contaminated soil DNA. Sets of approximately 1-kb DNA fragments were amplified with degenerate primers designed to amplify the C-terminal portion of bphA. These fragments were shuffled, and the resulting library was used to replace the corresponding fragment of Burkholderia xenovorans LB400 bphA. Variants were screened for their ability to oxygenate 2,2'-CB onto carbons 5 and 6, which are positions that LB400 BPDO is unable to attack. Variants S100, S149, and S151 were obtained and exhibited this feature. Variant S100 BPDO produced exclusively cis-5,6-dihydro-5,6-dihydroxy-2,2'-dichlorobiphenyl from 2,2'-CB. Moreover, unlike LB400 BPDO, S100 BphA catalyzed the oxygenation of 2,2',3,3'-tetrachlorobiphenyl onto carbons 5 and 6 exclusively and it was unable to oxygenate 2,2',5,5'-tetrachlorobiphenyl. Based on oxygen consumption measurements, variant S100 oxygenated 2,2'-CB at a rate of 16 +/- 1 nmol min(-1) per nmol enzyme, which was similar to the value observed for LB400 BPDO. cis-5,6-Dihydro-5,6-dihydroxy-2,2'-dichlorobiphenyl was further oxidized by 2,3-dihydro-2,3-dihydroxybiphenyl dehydrogenase (BphB) and 2,3-dihydroxybiphenyl dioxygenase (BphC). Variant S100 was, in addition, able to oxygenate benzene, toluene, and ethyl benzene. Sequence analysis identified amino acid residues M237 S238 and S283 outside regions III and IV that influence the activity toward doubly ortho-substituted chlorobiphenyls.  相似文献   

10.
11.
12.
The 80-kb dit cluster of Burkholderia xenovorans LB400 encodes the catabolism of abietane diterpenoids. This cluster includes ditQ and ditU, predicted to encode cytochromes P450 (P450s) belonging to the poorly characterized CYP226A subfamily. Using proteomics, we identified 16 dit-encoded proteins that were significantly more abundant in LB400 cells grown on dehydroabietic acid (DhA) or abietic acid (AbA) than in succinate-grown cells. A key difference in the catabolism of DhA and AbA lies in the differential expression of the P450s; DitU was detected only in the AbA-grown cells, whereas DitQ was expressed both during growth on DhA and during growth on AbA. Analyses of insertion mutants showed that ditQ was required for growth on DhA, ditU was required for growth on AbA, and neither gene was required for growth on the central intermediate, 7-oxo-DhA. In cell suspension assays, patterns of substrate removal and metabolite accumulation confirmed the role of DitU in AbA transformation and the role of DitQ in DhA transformation. Spectral assays revealed that DitQ binds both DhA (dissociation constant, 0.98 ± 0.01 μM) and palustric acid. Finally, DitQ transformed DhA to 7-hydroxy-DhA in vitro. These results demonstrate the distinct roles of the P450s DitQ and DitU in the transformation of DhA and AbA, respectively, to 7-oxo-DhA in a convergent degradation pathway.  相似文献   

13.
We examined the metabolism of dibenzofuran (DF) and dibenzo-p-dioxin (DD) by the biphenyl dioxygenase (BPDO) of Comamonas testosteroni B-356 and compared it with that of Burkholderia xenovorans LB400. Data showed that both enzymes oxygenated DF at a low rate, but Escherichia coli cells expressing LB400 BPDO degraded DF at higher rate (30 nmol in 18 h) compared with cells expressing B-356 BPDO (2 nmol in 18 h). Furthermore, both BPDOs produced dihydro-dihydroxy-dibenzofuran as a major metabolite, which resulted from the lateral oxygenation of DF. 2,2,3-Trihydroxybiphenyl (resulting from angular oxygenation of DF) was a minor metabolite produced by both enzymes. Deuterated DF was used to demonstrate the production of 2,2,3-dihydroxybiphenyl through angular oxygenation of DF. When tested for their ability to oxygenate DD, both enzymes produced as sole metabolite, 2,2,3-trihydroxybiphenyl ether at about the same rate, indicating similar catalytic properties toward this substrate. Altogether, although LB400 and B-356 BPDOs oxygenate a different range of chlorobiphenyls, their metabolite profiles toward DF and DD are similar. This suggests that co-planarity influences the regiospecificity of BPDO toward DF and DD to a higher extent than the presence of an ortho substituent on the molecule.  相似文献   

14.
2,2'-Dichlorobiphenyl (CB) is transformed by the biphenyl dioxygenase of Burkholderia xenovorans LB400 (LB400 BPDO) into two metabolites (1 and 2). The most abundant metabolite, 1, was previously identified as 2,3-dihydroxy-2'-chlorobiphenyl and was presumed to originate from the initial attack by the oxygenase on the chlorine-bearing ortho carbon and on its adjacent meta carbon of one phenyl ring. 2,3,2',3'-Tetrachlorobiphenyl is transformed by LB400 BPDO into two metabolites that had never been fully characterized structurally. We determined the precise identity of the metabolites produced by LB400 BPDO from 2,2'-CB and 2,3,2',3'-CB, thus providing new insights on the mechanism by which 2,2'-CB is dehalogenated to generate 2,3-dihydroxy-2'-chlorobiphenyl. We reacted 2,2'-CB with the BPDO variant p4, which produces a larger proportion of metabolite 2. The structure of this compound was determined as cis-3,4-dihydro-3,4-dihydroxy-2,2'-dichlorobiphenyl by NMR. Metabolite 1 obtained from 2,2'-CB-d(8) was determined to be a dihydroxychlorobiphenyl-d(7) by gas chromatographic-mass spectrometric analysis, and the observed loss of only one deuterium clearly shows that the oxygenase attack occurs on carbons 2 and 3. An alternative attack at the 5 and 6 carbons followed by a rearrangement leading to the loss of the ortho chlorine would have caused the loss of more than one deuterium. The major metabolite produced from catalytic oxygenation of 2,3,2',3'-CB by LB400 BPDO was identified by NMR as cis-4,5-dihydro-4,5-dihydroxy-2,3,2',3'-tetrachlorobiphenyl. These findings show that LB400 BPDO oxygenates 2,2'-CB principally on carbons 2 and 3 and that BPDO regiospecificity toward 2,2'-CB and 2,3,2,',3'-CB disfavors the dioxygenation of the chlorine-free ortho-meta carbons 5 and 6 for both congeners.  相似文献   

15.
The recently identified benzoate oxidation (box) pathway in Burkholderia xenovorans LB400 (LB400 hereinafter) assimilates benzoate through a unique mechanism where each intermediate is processed as a coenzyme A (CoA) thioester. A key step in this process is the conversion of 3,4-dehydroadipyl-CoA semialdehyde into its corresponding CoA acid by a novel aldehyde dehydrogenase (ALDH) (EC 1.2.1.x). The goal of this study is to characterize the biochemical and structural properties of the chromosomally encoded form of this new class of ALDHs from LB400 (ALDHC) in order to better understand its role in benzoate degradation. To this end, we carried out kinetic studies with six structurally diverse aldehydes and nicotinamide adenine dinucleotide (phosphate) (NAD + and NADP +). Our data definitively show that ALDHC is more active in the presence of NADP + and selective for linear medium-chain to long-chain aldehydes. To elucidate the structural basis for these biochemical observations, we solved the 1.6-Å crystal structure of ALDHC in complex with NADPH bound in the cofactor-binding pocket and an ordered fragment of a polyethylene glycol molecule bound in the substrate tunnel. These data show that cofactor selectivity is governed by a complex network of hydrogen bonds between the oxygen atoms of the 2′-phosphoryl moiety of NADP + and a threonine/lysine pair on ALDHC. The catalytic preference of ALDHC for linear longer-chain substrates is mediated by a deep narrow configuration of the substrate tunnel. Comparative analysis reveals that reorientation of an extended loop (Asn478-Pro490) in ALDHC induces the constricted structure of the substrate tunnel, with the side chain of Asn478 imposing steric restrictions on branched-chain and aromatic aldehydes. Furthermore, a key glycine (Gly104) positioned at the mouth of the tunnel allows for maximum tunnel depth required to bind medium-chain to long-chain aldehydes. This study provides the first integrated biochemical and structural characterization of a box-pathway-encoded ALDH from any organism and offers insight into the catalytic role of ALDHC in benzoate degradation.  相似文献   

16.
Recent microarray experiments suggested that Burkholderia xenovorans LB400, a potent polychlorinated biphenyl (PCB)-degrading bacterium, utilizes up to three apparently redundant benzoate pathways and a C(1) metabolic pathway during biphenyl and benzoate metabolism. To better characterize the roles of these pathways, we performed quantitative proteome profiling of cells grown on succinate, benzoate, or biphenyl and harvested during either mid-logarithmic growth or the transition between the logarithmic and stationary growth phases. The Bph enzymes, catabolizing biphenyl, were approximately 16-fold more abundant in biphenyl- versus succinate-grown cells. Moreover, the upper and lower bph pathways were independently regulated. Expression of each benzoate pathway depended on growth substrate and phase. Proteins specifying catabolism via benzoate dihydroxylation and catechol ortho-cleavage (ben-cat pathway) were approximately an order of magnitude more abundant in benzoate- versus biphenyl-grown cells at the same growth phase. The chromosomal copy of the benzoyl-coenzyme A (CoA) (box(C)) pathway was also expressed during growth on biphenyl: Box(C) proteins were approximately twice as abundant as Ben and Cat proteins under these conditions. By contrast, proteins of the megaplasmid copy of the benzoyl-CoA (box(M)) pathway were only detected in transition-phase benzoate-grown cells. Other proteins detected at increased levels in benzoate- and biphenyl-grown cells included general stress response proteins potentially induced by reactive oxygen species formed during aerobic aromatic catabolism. Finally, C(1) metabolic enzymes were present in biphenyl-grown cells during transition phase. This study provides insights into the physiological roles and integration of apparently redundant catabolic pathways in large-genome bacteria and establishes a basis for investigating the PCB-degrading abilities of this strain.  相似文献   

17.
It is now established that several amino acids of region III of the biphenyl dioxygenase (BPDO) alpha subunit are involved in substrate recognition and regiospecificity toward chlorobiphenyls. However, the sequence pattern of the amino acids of that segment of seven amino acids located in the C-terminal portion of the alpha subunit is rather limited in BPDOs of natural occurrence. In this work, we have randomly mutated simultaneously four residues (Thr(335)-Phe(336)-Ile(338)-Ile(341)) of region III of Burkholderia xenovorans LB400 BphA. The library was screened for variants able to oxygenate 2,2'-dichlorobiphenyl (2,2'-CB). Replacement of Phe(336) with Met or Ile with a concomitant change of Thr(335) to Ala created new variants that transformed 2,2'-CB into 3,4-dihydro-3,4-dihydroxy-2,2'-dichlorobiphenyl, which is a dead end metabolite that was not cleaved by BphC. Replacement of Thr(335)-Phe(336) with Ala(335)-Leu(336) did not cause this type of phenotypic change. Regiospecificity toward congeners other than 2,2'-CB that were oxygenated more efficiently by variant Ala(335)-Met(336) than by LB400 BPDO was similar for both enzymes. Thus structural changes that altered the regiospecificity toward 2,2'-CB did not affect the metabolite profile of other congeners, although it affected the rate of conversion of these congeners. It was especially noteworthy that both LB400 BPDO and the Ala(335)-Met(336) variant generated 2,3-dihydroxy-2',4,4'-trichlorobiphenyl as the sole metabolite from 2,4,2',4'-CB and 4,5-dihydro-4,5-dihydroxy-2,3,2',3'-tetrachlorobiphenyl as the major metabolite from 2,3,2',3'-CB. This shows that 2,4,2',4'-CB is oxygenated principally onto vicinal ortho-meta carbons 2 and 3 and that 2,3,2',3'-CB is oxygenated onto meta-para carbons 4 and 5 by both enzymes. The data suggest that interactions between the chlorine substitutes on the phenyl ring and specific amino acid residues of the protein influence the orientation of the phenyl ring inside the catalytic pocket.  相似文献   

18.
19.
BphD of Burkholderia xenovorans LB400 catalyzes an unusual C-C bond hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) to afford benzoic acid and 2-hydroxy-2,4-pentadienoic acid (HPD). An enol-keto tautomerization has been proposed to precede hydrolysis via a gem-diol intermediate. The role of the canonical catalytic triad (Ser-112, His-265, Asp-237) in mediating these two half-reactions remains unclear. We previously reported that the BphD-catalyzed hydrolysis of HOPDA (lambda(max) is 434 nm for the free enolate) proceeds via an unidentified intermediate with a red-shifted absorption spectrum (lambda(max) is 492 nm) (Horsman, G. P., Ke, J., Dai, S., Seah, S. Y. K., Bolin, J. T., and Eltis, L. D. (2006) Biochemistry 45, 11071-11086). Here we demonstrate that the S112A variant generates and traps a similar intermediate (lambda(max) is 506 nm) with a similar rate, 1/tau approximately 500 s(-1). The crystal structure of the S112A:HOPDA complex at 1.8-A resolution identified this intermediate as the keto tautomer, (E)-2,6-dioxo-6-phenyl-hex-3-enoate. This keto tautomer did not accumulate in either the H265A or the S112A/H265A double variants, indicating that His-265 catalyzes tautomerization. Consistent with this role, the wild type and S112A enzymes catalyzed tautomerization of the product HPD, whereas H265A variants did not. This study thus identifies a keto intermediate, and demonstrates that the catalytic triad histidine catalyzes the tautomerization half-reaction, expanding the role of this residue from its purely hydrolytic function in other serine hydrolases. Finally, the S112A:HOPDA crystal structure is more consistent with hydrolysis occurring via an acyl-enzyme intermediate than a gem-diol intermediate as solvent molecules have poor access to C6, and the closest ordered water is 7 A away.  相似文献   

20.
Horsman GP  Ke J  Dai S  Seah SY  Bolin JT  Eltis LD 《Biochemistry》2006,45(37):11071-11086
Kinetic and structural analyses of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) hydrolase from Burkholderia xenovorans LB400 (BphD(LB400)) provide insight into the catalytic mechanism of this unusual serine hydrolase. Single turnover stopped-flow analysis at 25 degrees C showed that the enzyme rapidly (1/tau(1) approximately 500 s(-1)) transforms HOPDA (lambda(max) = 434 nm) into a species with electronic absorption maxima at 473 and 492 nm. The absorbance of this enzyme-bound species (E:S) decayed in a biphasic manner (1/tau(2) = 54 s(-1), 1/tau(3) = 6 s(-1) approximately k(cat)) with simultaneous biphasic appearance (48 and 8 s(-1)) of an absorbance band at 270 nm characteristic of one of the products, 2-hydroxypenta-2,4-dienoic acid (HPD). Increasing solution viscosity with glycerol slowed 1/tau(1) and 1/tau(2) but affected neither 1/tau(3) nor k(cat), suggesting that 1/tau(2) may reflect diffusive HPD dissociation, and 1/tau(3) represents an intramolecular event. Product inhibition studies suggested that the other product, benzoate, is released after HPD. Contrary to studies in a related hydrolase, we found no evidence that ketonized HOPDA is partially released prior to hydrolysis, and, therefore, postulate that the biphasic kinetics reflect one of two mechanisms, pending assignment of E:S (lambda(max) = 492 nm). The crystal structures of the wild type, the S112C variant, and S112C incubated with HOPDA were each determined to 1.6 A resolution. The latter reveals interactions between conserved active site residues and the dienoate moiety of the substrate. Most notably, the catalytic residue His265 is hydrogen-bonded to the 2-hydroxy/oxo substituent of HOPDA, consistent with a role in catalyzing ketonization. The data are more consistent with an acyl-enzyme mechanism than with the formation of a gem-diol intermediate.  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