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 406 毫秒
1.
The ability of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate (HOPDA) hydrolase (BphD) of Burkholderia cepacia LB400 to hydrolyze polychlorinated biphenyl (PCB) metabolites was assessed by determining its specificity for monochlorinated HOPDAs. The relative specificities of BphD for HOPDAs bearing chlorine substituents on the phenyl moiety were 0.28, 0.38, and 1.1 for 8-Cl, 9-Cl, and 10-Cl HOPDA, respectively, versus HOPDA (100 mm phosphate, pH 7.5, 25 degrees C). In contrast, HOPDAs bearing chlorine substituents on the dienoate moiety were poor substrates for BphD, which hydrolyzed 3-Cl, 4-Cl, and 5-Cl HOPDA at relative maximal rates of 2.1 x 10(-3), 1.4 x 10(-4), and 0.36, respectively, versus HOPDA. The enzymatic transformation of 3-, 5-, 8-, 9-, and 10-Cl HOPDAs yielded stoichiometric quantities of the corresponding benzoate, indicating that BphD catalyzes the hydrolysis of these HOPDAs in the same manner as unchlorinated HOPDA. HOPDAs also underwent a nonenzymatic transformation to products that included acetophenone. In the case of 4-Cl HOPDA, this transformation proceeded via the formation of 4-OH HOPDA (t(12) = 2.8 h; 100 mm phosphate, pH 7.5, 25 degrees C). 3-Cl HOPDA (t(12) = 504 h) was almost 3 times more stable than 4-OH HOPDA. Finally, 3-Cl, 4-Cl and 4-OH HOPDAs competitively inhibited the BphD-catalyzed hydrolysis of HOPDA (K(ic) values of 0.57 +/- 0. 04, 3.6 +/- 0.2, and 0.95 +/- 0.04 microm, respectively). These results explain the accumulation of HOPDAs and chloroacetophenones in the microbial degradation of certain PCB congeners. More significantly, they indicate that in the degradation of PCB mixtures, BphD would be inhibited, thereby slowing the mineralization of all congeners. BphD is thus a key determinant in the aerobic microbial degradation of PCBs.  相似文献   

2.
2-Hydroxy-6-oxo-6-phenylhexa-2,4-dienoate (HOPDA) hydrolase (BphD) is a key determinant in the aerobic transformation of polychlorinated biphenyls (PCBs) by Burkholderia sp. strain LB400 (S. Y. K. Seah, G. Labbé, S. Nerdinger, M. Johnson, V. Snieckus, and L. D. Eltis, J. Biol. Chem. 275:15701-15708, 2000). To determine whether this is also true in divergent biphenyl degraders, the homologous hydrolase of Rhodococcus globerulus P6, BphD(P6), was hyperexpressed, purified to apparent homogeneity, and studied by steady-state kinetics. BphD(P6) hydrolyzed HOPDA with a k(cat)/K(m) of 1.62 (+/- 0.03) x 10(7) M(-1) s(-1) (100 mM phosphate [pH 7.5], 25 degrees C), which is within 70% of that of BphD(LB400). BphD(P6) was also similar to BphD(LB400) in that it catalyzed the hydrolysis of HOPDAs bearing chloro substituents on the phenyl moiety at least 25 times more specifically than those bearing chloro substituents on the dienoate moiety. However, the rhodococcal enzyme was significantly more specific for 9-Cl and 10-Cl HOPDAs, catalyzing the hydrolysis of 9-Cl, 10-Cl, and 9,10-diCl HOPDAs two- to threefold respectively, more specifically than HOPDA. Moreover, 4-Cl HOPDA competitively inhibited BphD(P6) more effectively than 3-Cl HOPDA, which is the inverse of what was observed in BphD(LB400). These results demonstrate that BphD is a key determinant in the aerobic transformation of PCBs by divergent biphenyl degraders, but that there exists significant diversity in the specificity of these biphenyl hydrolases.  相似文献   

3.
BphK is a glutathione S-transferase of unclear physiological function that occurs in some bacterial biphenyl catabolic (bph) pathways. We demonstrated that BphK of Burkholderia xenovorans strain LB400 catalyzes the dehalogenation of 3-chloro 2-hydroxy-6-oxo-6-phenyl-2,4-dienoates (HOPDAs), compounds that are produced by the cometabolism of polychlorinated biphenyls (PCBs) by the bph pathway and that inhibit the pathway's hydrolase. A one-column protocol was developed to purify heterologously produced BphK. The purified enzyme had the greatest specificity for 3-Cl HOPDA (kcat/Km, approximately 10(4) M(-1) s(-1)), which it dechlorinated approximately 3 orders of magnitude more efficiently than 4-chlorobenzoate, a previously proposed substrate of BphK. The enzyme also catalyzed the dechlorination of 5-Cl HOPDA and 3,9,11-triCl HOPDA. By contrast, BphK did not detectably transform HOPDA, 4-Cl HOPDA, or chlorinated 2,3-dihydroxybiphenyls. The BphK-catalyzed dehalogenation proceeded via a ternary-complex mechanism and consumed 2 equivalents of glutathione (GSH) (Km for GSH in the presence of 3-Cl HOPDA, approximately 0.1 mM). A reaction mechanism consistent with the enzyme's specificity is proposed. The ability of BphK to dehalogenate inhibitory PCB metabolites supports the hypothesis that this enzyme was recruited to facilitate PCB degradation by the bph pathway.  相似文献   

4.
The microbial degradation of polychlorinated biphenyls (PCBs) by the biphenyl catabolic (Bph) pathway is limited in part by the pathway's fourth enzyme, BphD. BphD catalyzes an unusual carbon-carbon bond hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA), in which the substrate is subject to histidine-mediated enol-keto tautomerization prior to hydrolysis. Chlorinated HOPDAs such as 3-Cl HOPDA inhibit BphD. Here we report that BphD preferentially hydrolyzed a series of 3-substituted HOPDAs in the order H>F>Cl>Me, suggesting that catalysis is affected by steric, not electronic, determinants. Transient state kinetic studies performed using wild-type BphD and the hydrolysis-defective S112A variant indicated that large 3-substituents inhibited His-265-catalyzed tautomerization by 5 orders of magnitude. Structural analyses of S112A.3-Cl HOPDA and S112A.3,10-diF HOPDA complexes revealed a non-productive binding mode in which the plane defined by the carbon atoms of the dienoate moiety of HOPDA is nearly orthogonal to that of the proposed keto tautomer observed in the S112A.HOPDA complex. Moreover, in the 3-Cl HOPDA complex, the 2-hydroxo group is moved by 3.6 A from its position near the catalytic His-265 to hydrogen bond with Arg-190 and access of His-265 is blocked by the 3-Cl substituent. Nonproductive binding may be stabilized by interactions involving the 3-substituent with non-polar side chains. Solvent molecules have poor access to C6 in the S112A.3-Cl HOPDA structure, more consistent with hydrolysis occurring via an acyl-enzyme than a gem-diol intermediate. These results provide insight into engineering BphD for PCB degradation.  相似文献   

5.
Ruzzini AC  Horsman GP  Eltis LD 《Biochemistry》2012,51(29):5831-5840
meta-Cleavage product (MCP) hydrolases catalyze C-C bond fission in the aerobic catabolism of aromatic compounds by bacteria. These enzymes utilize a Ser-His-Asp triad to catalyze hydrolysis via an acyl-enzyme intermediate. BphD, which catalyzes the hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) in biphenyl degradation, catalyzed the hydrolysis of an ester analogue, p-nitrophenyl benzoate (pNPB), with a k(cat) value (6.3 ± 0.5 s(-1)) similar to that of HOPDA (6.5 ± 0.5 s(-1)). Consistent with the breakdown of a shared intermediate, product analyses revealed that BphD catalyzed the methanolysis of both HOPDA and pNPB, partitioning the products to benzoic acid and methyl benzoate in similar ratios. Turnover of HOPDA was accelerated up to 4-fold in the presence of short, primary alcohols (methanol > ethanol > n-propanol), suggesting that deacylation is rate-limiting during catalysis. In the steady-state hydrolysis of HOPDA, k(cat)/K(m) values were independent of methanol concentration, while both k(cat) and K(m) values increased with methanol concentration. This result was consistent with a simple model of nucleophilic catalysis. Although the enzyme could not be saturated with pNPB at methanol concentrations of >250 mM, k(obs) values from the steady-state turnover of pNPB at low methanol concentrations were also consistent with a nucleophilic mechanism of catalysis. Finally, transient-state kinetic analysis of pNPB hydrolysis by BphD variants established that substitution of the catalytic His reduced the rate of acylation by more than 3 orders of magnitude. This suggests that for pNPB hydrolysis, the serine nucleophile is activated by the His-Asp dyad. In contrast, rapid acylation of the H265Q variant during C-C bond cleavage suggests that the serinate forms via a substrate-assisted mechanism. Overall, the data indicate that ester hydrolysis proceeds via the same acyl-enzyme intermediate as that of the physiological substrate but that the serine nucleophile is activated via a different mechanism.  相似文献   

6.
Prokaryotic glutathione S-transferases are as diverse as their eukaryotic counterparts but are much less well characterized. BphK from Burkholderia xenovorans LB400 consumes two GSH molecules to reductively dehalogenate chlorinated 2-hydroxy-6-oxo-6-phenyl-2,4-dienoates (HOPDAs), inhibitory polychlorinated biphenyl metabolites. Crystallographic structures of two ternary complexes of BphK were solved to a resolution of 2.1A. In the BphK-GSH-HOPDA complex, GSH and HOPDA molecules occupy the G- and H-subsites, respectively. The thiol nucleophile of the GSH molecule is positioned for SN2 attack at carbon 3 of the bound HOPDA. The respective sulfur atoms of conserved Cys-10 and the bound GSH are within 3.0A, consistent with product release and the formation of a mixed disulfide intermediate. In the BphK-(GSH)2 complex, a GSH molecule occupies each of the two subsites. The three sulfur atoms of the two GSH molecules and Cys-10 are aligned suitably for a disulfide exchange reaction that would regenerate the resting enzyme and yield disulfide-linked GSH molecules. A second conserved residue, His-106, is adjacent to the thiols of Cys-10 and the GSH bound to the G-subsite and thus may stabilize a transition state in the disulfide exchange reaction. Overall, the structures support and elaborate a proposed dehalogenation mechanism for BphK and provide insight into the plasticity of the H-subsite.  相似文献   

7.
Recent studies demonstrated that 2,3-dihydroxybiphenyl 1,2-dioxygenase from Burkholderia sp. strain LB400 (DHBDLB400; EC 1.13.11.39) cleaves chlorinated 2,3-dihydroxybiphenyls (DHBs) less specifically than unchlorinated DHB and is competitively inhibited by 2',6'-dichloro-2,3-dihydroxybiphenyl (2',6'-diCl DHB). To determine whether these are general characteristics of DHBDs, we characterized DHBDP6-I and DHBDP6-III, two evolutionarily divergent isozymes from Rhodococcus globerulus strain P6, another good polychlorinated biphenyl (PCB) degrader. In contrast to DHBDLB400, both rhodococcal enzymes had higher specificities for some chlorinated DHBs in air-saturated buffer. Thus, DHBDP6-I cleaved the DHBs in the following order of specificity: 6-Cl DHB > 3'-Cl DHB approximately DHB approximately 4'-Cl DHB > 2'-Cl DHB > 4-Cl DHB > 5-Cl DHB. It also cleaved its preferred substrate, 6-Cl DHB, three times more specifically than DHB. Interestingly, some of the worst substrates for DHBDP6-I were among the best for DHBDP6-III (4-Cl DHB > 5-Cl DHB approximately 6-Cl DHB approximately 3'-Cl DHB > DHB > 2'-Cl DHB approximately 4'-Cl DHB; DHBDP6-III cleaved 4-Cl DHB two times more specifically than DHB). Generally, each of the monochlorinated DHBs inactivated the enzymes more rapidly than DHB. The exceptions were 4-Cl DHB for DHBDP6-I and 2'-Cl DHB for DHBDP6-III. As observed in DHBDLB400, chloro substituents influenced the reactivity of the dioxygenases with O2. For example, the apparent specificities of DHBDP6-I and DHBDP6-III for O2 in the presence of 2'-Cl DHB were lower than those in the presence of DHB by factors of >60 and 4, respectively. DHBDP6-I and DHBDP6-III shared the relative inability of DHBDLB400 to cleave 2',6'-diCl DHB (apparent catalytic constants of 0.088 +/- 0.004 and 0.069 +/- 0.002 s(-1), respectively). However, these isozymes had remarkably different apparent K(m) values for this compound (0.007 +/- 0.001, 0.14 +/- 0.01, and 3.9 +/- 0.4 micro M for DHBDLB400, DHBDP6-I, and DHBDP6-III, respectively). The markedly different reactivities of DHBDP6-I and DHBDP6-III with chlorinated DHBs undoubtedly contribute to the PCB-degrading activity of R. globerulus P6.  相似文献   

8.
Hydrolysis following meta-ring cleavage by a dioxygenase is a well-known step in aromatic compound metabolism. The 2-hydroxy-6-oxo-6-(2'-aminophenyl)hexa-2,4-dienoic acid hydrolase from Pseudomonas LD2 is a new member of the small group of characterized aromatic hydrolases that catalyze the cleavage of C-C bonds. In this study, the His(6)-tagged 2-hydroxy-6-oxo-6-(2'-aminophenyl)hexa-2,4-dienoic acid (HOPDA) hydrolase was purified from a recombinant Escherichia coli strain utilizing immobilized metal affinity chromatography. 2-Hydroxy-6-oxo-6-(2'-aminophenyl)hexa-2,4-dienoic acid hydrolase is a colorless homodimer with no cofactor requirement. The enzyme actively converted HOPDA into benzoic acid and 2-hydroxypenta-2,4-dienoic acid. The enzyme exhibited activity between pH 6.5 and 10.5 with a maximum activity at pH 7.0. The optimum temperature at pH 7.0 was 60 degrees C. The calculated K'(m) for HOPDA was 4.6 microM, the V(max) was 3.3 micromol min(-1), and the K(s) was 70.0 microM. This corresponds to a maximum specific turnover rate of 1300 HOPDAs(-1)dimer(-1). The deduced amino acid sequence of CarC showed 30.3, 31.3, and 31.8% identity with TodF (P. putida F1), XylF (P. putida), and DmpD (Pseudomonas sp. CF600), respectively, which are meta-cleavage compound hydrolases from other Pseudomonads. The amino acid sequence Gly-X-Ser-X-Gly, which is highly conserved in these hydrolases, is also found in CarC. Lysates from a strain expressing enzyme in which the putative active site serine is mutated to alanine showed a significant reduction in activity.  相似文献   

9.
10.
In the recently identified cholesterol catabolic pathway of Mycobacterium tuberculosis, 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate hydrolase (HsaD) is proposed to catalyze the hydrolysis of a carbon-carbon bond in 4,5–9,10-diseco-3-hydroxy-5,9,17-tri-oxoandrosta-1(10),2-diene-4-oic acid (DSHA), the cholesterol meta-cleavage product (MCP) and has been implicated in the intracellular survival of the pathogen. Herein, purified HsaD demonstrated 4–33 times higher specificity for DSHA (kcat/Km = 3.3 ± 0.3 × 104 m−1 s−1) than for the biphenyl MCP 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) and the synthetic analogue 8-(2-chlorophenyl)-2-hydroxy-5-methyl-6-oxoocta-2,4-dienoic acid (HOPODA), respectively. The S114A variant of HsaD, in which the active site serine was substituted with alanine, was catalytically impaired and bound DSHA with a Kd of 51 ± 2 μm. The S114A·DSHA species absorbed maximally at 456 nm, 60 nm red-shifted versus the DSHA enolate. Crystal structures of the variant in complex with HOPDA, HOPODA, or DSHA to 1.8–1.9 Åindicate that this shift is due to the enzyme-induced strain of the enolate. These data indicate that the catalytic serine catalyzes tautomerization. A second role for this residue is suggested by a solvent molecule whose position in all structures is consistent with its activation by the serine for the nucleophilic attack of the substrate. Finally, the α-helical lid covering the active site displayed a ligand-dependent conformational change involving differences in side chain carbon positions of up to 6.7 Å, supporting a two-conformation enzymatic mechanism. Overall, these results provide novel insights into the determinants of specificity in a mycobacterial cholesterol-degrading enzyme as well as into the mechanism of MCP hydrolases.  相似文献   

11.
meta-Cleavage product (MCP) hydrolases (EC 3.7.1.9) can catalyze a specific C–C bond fission during the microbial aerobic degradation of aromatics. The previous studies on structure–function relationship of MCP hydrolases mainly focus on the active site residues by site-directed mutagenesis. However, the information about the role of the non-active-site residues is still unclear. In this study, a non-active-site residue Met148 of MCP hydrolase BphD was selected as the mutagenesis site according to the sequence alignments, structure superimpose and the tunnel analysis, which underwent the saturation mutagenesis resulting 19 mutants. The catalytic efficiencies of the mutants on 6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) were all decreased compared with the wild-type one except for the M148D mutant. Especially, the M148P mutant exhibited 290-fold lower k cat/K m than that of the wild-type BphD. Transient kinetic analyses of M148P showed the reciprocal relaxation time corresponded to C–C bond cleavage and product release steps (9.6 s?1) was 4.08-fold lower than BphD WT (39.2 s?1). Tunnel cluster analysis of BphD WT, M148P and M148W demonstrated that only the bulky Trp148 could block tunnel T2 in the BphD WT, but it exhibited slight effects on the catalytic efficiency (0.94-fold of BphD WT). Therefore, product release was not the main reason for the efficiency decrease of M148P. On the other hand, molecular dynamics simulations on the BphD WT and BphD M148P in complex with HOPDA indicated that the dramatic decrease of the catalytic efficiencies of BphD M148P should be due to the unproductive binding of HOPDA. The study demonstrated the catalytic efficiency of MCP hydrolase can be engineered by modification of non-active site residue.  相似文献   

12.
The double bonds of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) were stabilized by methylation to establish which of the double bonds of the meta ring-fission compound of biphenyl was reduced by the HOPDA reducing enzyme. HOPDA reducing enzyme III converted 2-methoxy-6-oxo-6-phenylhexa-2,4-dienoic acid methyl ester into 2-methoxy-6-oxo-6-phenylhexa-2-enoic acid methyl ester. To discover the metabolic pathway of HOPDA, partially purified enzyme fractions were used. The eluate from a 2nd column of DEAE-cellulose transformed HOPDA to γ-benzoylbutyric acid, 2,6-dioxo-6-phenylhexanoic acid, and γ-benzoylbutyraldehyde. Fractions passed through the 1st column of DEAE-cellulose formed γ-benzoylbutyric acid and 2-hydroxy-6-oxo-6-phenylhexanoic acid from HOPDA. Based on these data and previous reports, a new metabolic divergence of biphenyl and related compounds was proposed.  相似文献   

13.
Oxidative damage to DNA has been reported to occur in a wide variety of disease states. The most widely used "marker" for oxidative DNA damage is 8-hydroxyguanine. However, the use of only one marker has limitations. Exposure of calf thymus DNA to an .OH-generating system (CuCl(2), ascorbate, H(2)O(2)) or to hypochlorous acid (HOCl), led to the extensive production of multiple oxidized or chlorinated DNA base products, as measured by gas chromatography-mass spectrometry. The addition of peroxynitrite (ONOO(-)) (<200 microM) or SIN-1 (1mM) to oxidized DNA led to the extensive loss of 8-hydroxyguanine, 5-hydroxycytosine, 2,6-diamino-4-hydroxy-5-formamidopyrimidine, 2-hydroxyadenine, 8-hydroxyadenine, and 4,6-diamino-5-formamidopyrimidine were lost at higher ONOO(-) concentrations (>200 microM). Exposure of DNA to HOCl led to the generation of 5-Cl uracil and 8-Cl adenine and addition of ONOO(-) (<200 microM) or SIN-1 (1mM) led to an extensive loss of 8-Cl adenine and a small loss of 5-Cl uracil at higher concentrations (>500 microM). An .OH-generating system (CuCl(2)/ascorbate/H(2)O(2)) could also destroy these chlorinated species. Treatment of oxidized or chlorinated DNA with acidified nitrite (NO(2)(-), pH 3) led to substantial loss of various base lesions, in particular 8-OH guanine, 5-OH cytosine, thymine glycol, and 8-Cl adenine. Our data indicate the possibility that when ONOO(-), nitrite in regions of low pH or .OH are produced at sites of inflammation, levels of certain damaged DNA bases could represent an underestimate of ongoing DNA damage. This study emphasizes the need to examine more than one modified DNA base when assessing the role of reactive species in human disease.  相似文献   

14.
Pseudomonas paucimobilis Q1 originally isolated as biphenyl degrading organism (Furukawa et al. 1983), was shown to grow with naphthalene. After growth with biphenyl or naphthalene the strain synthesized the same enzyme for the ring cleavage of 2,3-dihydroxybiphenyl or 1,2-dihydroxynaphthalene. The enzyme, although characterized as 2,3-dihydroxybiphenyl dioxygenase (Taira et al. 1988), exhibited considerably higher relative activity with 1,2-dihydroxynaphthalene. These results demonstrate that this enzyme can function both in the naphthalene and biphenyl degradative pathway.Abbreviations DHBP dihydroxybiphenyl - DHBPDO 2,3-dihydroxybiphenyl dioxygenase - DHDHNDH 1,2-dihydroxy-1,2-dihydronaphthalene dehydrogenase - DHN 1,2-dihydroxynaphthalene - DHNDO 1,2-dihydroxynaphthalene dioxygenase - HBP cis-2-hydroxybenzalpyruvate - HOPDA 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate - PCB polychlorinated biphenyl - 2NS naphthalene-2-sulfonic acid  相似文献   

15.
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   

16.
The oxygenase component of biphenyl dioxygenase (BPDO) from Comamonas testosteroni B-356 dihydroxylates biphenyl and some polychlorinated biphenyls (PCBs), thereby initiating their degradation. Overexpressed, anaerobically purified BPDO had a specific activity of 4.9 units/mg, and its oxygenase component appeared to contain a full complement of Fe(2)S(2) center and catalytic iron. Oxygenase crystals in space group R3 were obtained under anaerobic conditions using polyethylene glycol as the precipitant. X-ray diffraction was measured to 1.6 A. Steady-state kinetics assays demonstrated that BPDO had an apparent k(cat)/K(m) for biphenyl of (1.2 +/- 0.1) x 10(6) M(-1) s(-1) in air-saturated buffer. Moreover, BPDO transformed dichlorobiphenyls (diClBs) in the following order of apparent specificities: 3,3'- > 2,2'- > 4, 4'-diClB. Strikingly, the ability of BPDO to utilize O(2) depended strongly on the biphenyl substrate: k(cat)/K(m(O(2))) = (3.6 +/- 0. 3), (0.06 +/- 0.02), and (0.4 +/- 0.07) x 10(5) M(-1) s(-1) in the presence of biphenyl and 2,2'- and 3,3'-diClBs, respectively. Moreover, biphenyl/O(2) consumed was 0.97, 0.44, 0.63, and 0.48 in the presence of biphenyl and 2,2'-, 3,3'-, and 4,4'-diClBs, respectively. Within experimental error, the balance of consumed O(2) was detected as H(2)O(2). Thus, PCB congeners such as 2, 2'-diClB exact a high energetic cost, produce a cytotoxic compound (H(2)O(2)), and can inhibit degradation of other congeners. Each of these effects would be predicted to inhibit the aerobic microbial catabolism of PCBs.  相似文献   

17.
We sought to determine whether microorganisms from the polychlorinated biphenyl (PCB)-contaminated sediment in Woods Pond (Lenox, Mass.) could dehalogenate brominated biphenyls. The PCB dechlorination specificities for the microorganisms in this sediment have been well characterized. This allowed us to compare the dehalogenation specificities for brominated biphenyls and chlorinated biphenyls within a single sediment. Anaerobic sediment microcosms were incubated separately at 25°C with 16 different mono- to tetrabrominated biphenyls (350 μM) and disodium malate (10 mM). Samples were extracted and analyzed by gas chromatography with an electron capture detector and a mass spectrometer detector at various times for up to 54 weeks. All of the tested brominated biphenyls were dehalogenated. For most congeners, including 2,6-dibromobiphenyl (26-BB) and 24-25-BB, the dehalogenation began within 1 to 2 weeks. However, for 246-BB and 2-2-BB, debromination was first observed at 7 and 14 weeks, respectively. Most intermediate products did not persist, but when 2-2-BB was produced as a dehalogenation product, it persisted for at least 15 weeks before it was dehalogenated to 2-BB and then to biphenyl. The dehalogenation specificities for brominated and chlorinated biphenyls were similar: meta and para substituents were generally removed first, and ortho substituents were more recalcitrant. However, the brominated biphenyls were better dehalogenation substrates than the chlorinated biphenyls. All of the tested bromobiphenyls, including those with ortho and unflanked meta and para substituents, were ultimately dehalogenated to biphenyl, whereas their chlorinated counterparts either were not dehalogenation substrates or were only partially dehalogenated. Our data suggest that PCB-dechlorinating microorganisms may be able to dehalogenate brominated biphenyls and may exhibit a relaxed specificity for these substrates.  相似文献   

18.
Tuning biphenyl dioxygenase for extended substrate specificity.   总被引:12,自引:0,他引:12  
Highly substituted polychlorinated biphenyls (PCBs) are known to be very resistant to aerobic biodegradation, particularly the initial attack by biphenyl dioxygenase. Functional evolution of the substrate specificity of biphenyl dioxygenase was demonstrated by DNA shuffling and staggered extension process (StEP) of the bphA gene coding for the large subunit of biphenyl dioxygenase. Several variants with an extended substrate range for PCBs were selected. In contrast to the parental biphenyl dioxygenases from Burkholderia cepacia LB400 and Pseudomonas pseudoalcaligenes KF707, which preferentially recognize either ortho- (LB400) or para- (KF707) substituted PCBs, several variants degraded both congeners to about the same extent. These variants also exhibited superior degradation capabilities toward several tetra- and pentachlorinated PCBs as well as commercial PCB mixtures, such as Aroclor 1242 or Aroclor 1254. Sequence analysis confirmed that most variants contained at least four to six template switches. All desired variants contained the Thr335Ala and Phe336Ile substitutions confirming the importance of this critical region in substrate specificity. These results suggest that the block-exchange nature of gene shuffling between a diverse class of dioxygenases may be the most useful approach for breeding novel dioxygenases for PCB degradation in the desired direction.  相似文献   

19.
The primary specificity residue of a substrate or an inhibitor, called the P(1) residue, is responsible for the proper recognition by the cognate enzyme. This residue enters the S(1) pocket of the enzyme and establishes contacts (up to 50%) inside the proteinase substrate cavity, strongly affecting its specificity. To analyze the influence on bovine alpha-chymotrypsin substrate activity, aromatic non-proteinogenic amino acid residues in position P(1) with the sequence Ac-Phe-Ala-Thr-X-Anb(5,2)-NH(2) were introduced: L-pyridyl alanine (Pal), 4-nitrophenylalanine - Phe(p-NO(2)), 4-aminophenylalanine - Phe(p-NH(2)), 4-carboxyphenylalanine Phe(p-COOH), 4-guanidine phenylalanine - Phe(p-guanidine), 4-methyloxycarbonyl-phenylalanine - Phe(p-COOMe), 4-cyanophenylalanine - Phe(p-CN), Phe, Tyr. The effect of the additional substituent at the phenyl ring of the Phe residue was investigated. All peptides contained an amide of 5-amino-2-nitrobenzoic acid, which served as a chromophore. Kinetic parameters (k(cat), K(M) and k(cat)/K(M)) of the peptides synthesized with bovine alpha-chymotrypsin were determined. The highest value of the specificity constant k(cat)/K(M), reaching 6.0 x 10(5) [M(-1)xs(-1)], was obtained for Ac-Phe-Ala-Thr-Phe(p-NO(2))-Anb(5,2)-NH(2). The replacement of the acetyl group with benzyloxycarbonyl moiety yielded a substrate with the value of k(cat) more than three times higher. Peptide aldehydes were synthesized with selected residues (Phe, Pal, Tyr, Phe(p-NO(2)) in position P(1) and potent chymotrypsin inhibitors were obtained. The dissociation constant (K(i)) with the experimental enzyme determined for the most active peptide, Tos-Phe-Ala-Thr-Phe(p-NO(2))-CHO, amounted to 1.12 x 10(-8) M.  相似文献   

20.
Hypochlorous acid (HOCl) is formed by the action of the enzyme myeloperoxidase on hydrogen peroxide and chloride ions. It has been shown to be highly bactericidal and cytotoxic by a variety of mechanisms, one of which, may be the modification of DNA. Previously we have demonstrated by GC-MS analysis that exposure of calf thymus DNA to HOCl causes extensive pyrimidine modification, including 5-chlorocytosine formation. Using GC-MS analysis, we now demonstrate the formation of an additional chlorinated base product, 8-Cl adenine. The addition of 50 μM HOCl was sufficient to produce a significant increase in this product. The reaction of HOCl with adenine in calf thymus DNA was shown to be rapid with the reaction complete after 1 min. pH-dependence studies suggest HOCl rather than its conjugate base (OCl-) to be responsible for 8-Cl adenine formation. Other commercially available chlorinated base products, 6-Cl guanine or 2-Cl adenine were not detected. Therefore, 8-Cl adenine might prove a useful biomarker for studying the role of reactive chlorine species (RCS) during inflammatory processes.  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