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.
A rare Gly210 deletion in protoporphyrinogen oxidase (PPO) was recently discovered in herbicide-resistant Amaranthus tuberculatus. According to the published X-ray structure of Nicotiana tabacum PPO, Gly210 is adjacent to, not in, the PPO active site, so it is a matter of interest to determine why its deletion imparts resistance to herbicides. In our kinetic experiments, this deletion did not affect the affinity of protoporphyrinogen IX nor the FAD content, but decreased the catalytic efficiency of the enzyme. The suboptimal Kcat was compensated by a significant increase in the Kis for inhibitors and a switch in their interactions from competitive to mixed-type inhibition. In our protein modeling studies on herbicide-susceptible PPO and resistant PPO, we show that Gly210 plays a key role in the αL helix-capping motif at the C-terminus of the α-8 helix which helps to stabilize the helix. In molecular dynamics simulations, the deletion had significant architecture consequences, destabilizing the α-8 helix-capping region and unraveling the last turn of the helix, leading to enlargement of the active site cavity by ∼ 50%. This seemingly innocuous deletion of Gly210 of the mitochondrial PPO imparts herbicide resistance to this dual-targeted protein without severely affecting its normal physiological function, which may explain why this unusual mutation was the favored evolutionary path for achieving resistance to PPO inhibitors.  相似文献   

2.
Tan Y  Sun L  Xi Z  Yang GF  Jiang DQ  Yan XP  Yang X  Li HY 《Analytical biochemistry》2008,383(2):200-204
Protoporphyrinogen oxidase (PPO) is a flavin adenine dinucleotide (FAD)-containing enzyme in the tetrapyrrole biosynthetic pathway that leads to the formation of both heme and chlorophylls, which has been identified as one of the most important action targets of commercial herbicides. The literature reports gave different PPO-catalytic kinetic parameters for the substrate protoporphyrinogen IX (Km of 0.1 to 10.4 μM) with different sources of PPO using fluorescent or HPLC methods. Herein we assayed the enzymatic activity of recombinant Bacillus subtilis PPO by using capillary electrophoresis (CE), a method with high separation efficiency, easy automation, and low sample consumption. The Michaelis constant and maximum reaction velocity were determined as 7.0 ± 0.6 μM and 0.38 ± 0.02 μmol min-1 μg−1, respectively. The interaction between PPO and acifluorfen, a commercial PPO-inhibiting herbicide, was measured as the inhibition constant 186.9 ± 9.3 μМ. The relationship between cofactor FAD and PPO activity can also be quantitatively studied by this CE method. The CE method used here should also be a convenient, reliable method for PPO study.  相似文献   

3.
In this study, a series of fluorine‐containing chiral hydrazide‐hydrazone derivatives [III‐XII] from ?‐cysteine ethyl ester hydrochloride was synthesized as new antioxidant and anticholinesterase agents. The antioxidant activity of these derivatives was evaluated by ABTS and DPPH· scavenging and CUPRAC assays and the anticholinesterase activity by the Ellman method spectrophotometrically. The results of the antioxidant assay showed that compounds V , IX , and X exhibited higher activity than BHT and α‐tocopherol used as positive standards. Among the synthesized derivatives, compound IX (IC50: 2.3 ± 1.6 μM) exhibited higher acetylcholinesterase inhibitory activity than galantamine (IC50: 4.5 ± 0.8 μM). Compounds XI (IC50: 9.6 ± 1.0 μM), IX (IC50: 12.5 ± 1.6 μM), III (IC50: 16.0 ± 1.6 μM), X (IC50: 17.2 ± 1.8 μM), VI (IC50: 20.2 ± 0.8 μM), XII (IC50: 21.5 ± 1.0 μM), and VII (IC50: 24.6 ± 0.6 μM) displayed better butyrylcholinesterase inhibitory activity than galantamine (IC50: 46.03 ± 0.14 μM). ADME‐Tox analysis was used to probe the drug‐like properties of the compounds. Molecular docking studies were also applied to understand the interactions between compounds and targets. The docking calculations were supported by the experimental data. In particular, compound IX , having better activity than galantamine against acetylcholinesterase and butyrylcholinesterase enzymes, was visualized using molecular docking.  相似文献   

4.
Chloramphenicol is a halogenated natural product bearing an unusual dichloroacetyl moiety that is critical for its antibiotic activity. The operon for chloramphenicol biosynthesis in Streptomyces venezuelae encodes the chloramphenicol halogenase CmlS, which belongs to the large and diverse family of flavin-dependent halogenases (FDH’s). CmlS was previously shown to be essential for the formation of the dichloroacetyl group. Here we report the X-ray crystal structure of CmlS determined at 2.2 Å resolution, revealing a flavin monooxygenase domain shared by all FDHs, but also a unique ‘winged-helix’ C-terminal domain that creates a T-shaped tunnel leading to the halogenation active site. Intriguingly, the C-terminal tail of this domain blocks access to the halogenation active site, suggesting a structurally dynamic role during catalysis. The halogenation active site is notably nonpolar and shares nearly identical residues with Chondromyces crocatus tyrosyl halogenase (CndH), including the conserved Lys (K71) that forms the reactive chloramine intermediate. The exception is Y350, which could be used to stabilize enolate formation during substrate halogenation. The strictly conserved residue E44, located near the isoalloxazine ring of the bound flavin adenine dinucleotide (FAD) cofactor, is optimally positioned to function as a remote general acid, through a water-mediated proton relay, which could accelerate the reaction of the chloramine intermediate during substrate halogenation, or the oxidation of chloride by the FAD(C4α)-OOH intermediate. Strikingly, the 8α carbon of the FAD cofactor is observed to be covalently attached to D277 of CmlS, a residue that is highly conserved in the FDH family. In addition to representing a new type of flavin modification, this has intriguing implications for the mechanism of FDHs. Based on the crystal structure and in analogy to known halogenases, we propose a reaction mechanism for CmlS.  相似文献   

5.
Sah S  Phale PS 《Biodegradation》2011,22(3):517-526
1-Naphthol 2-hydroxylase (1-NH) which catalyzes the conversion of 1-naphthol to 1,2-dihydroxynaphthalene was purified to homogeneity from carbaryl-degrading Pseudomonas sp. strain C6. The enzyme was found to be a homodimer with subunit molecular weight of 66 kDa. UV, visible and fluorescence spectral properties, identification of flavin moiety by HPLC as FAD, and reconstitution of apoenzyme by FAD suggest that enzyme is FAD-dependent. 1-NH accepts electron from NADH as well as NADPH. Besides 1-naphthol (K m, 9.1 μM), the enzyme also accepts 5-amino 1-naphthol (K m, 6.4 μM) and 4-chloro 1-naphthol (K m, 2.3 μM) as substrates. Enzyme showed substrate inhibition phenomenon at high concentration of 1-naphthol (K i, 283 μM). Stoichiometric consumption of oxygen and NADH, and biochemical properties suggest that 1-NH belongs to FAD containing external flavomonooxygenase group of oxido-reductase class of enzymes. Based on biochemical and kinetic properties, 1-NH from Pseudomonas sp. strain C6 appears to be different than that reported earlier from Pseudomonas sp. strain C4. Chemical modification and protection by 1-naphthol and NADH suggest that His, Arg, Cys, Tyr and Trp are at or near the active site of 1-NH.  相似文献   

6.
A mutant strain of Pyropia yezoensis, strain E, was isolated from the free‐living conchocelis of a pure strain (NA) treated with ethyl methane sulfonate. The incremental quantities of young strain E blades were higher than those of NA after 14 d of cultivation, indicating that young blades of mutant strain E released more archeospores. The mean length and weight of large E blades were both over three times greater than those of NA after 4 weeks of cultivation. The photosynthetic parameters (Fv/Fm, Y[I], Y[II], and O2 evolution rate) and pigment contents (including phycoerythrin and phycocyanin) of strain E blades were higher than those of NA (P < 0.05). The cellular respiratory rate of strain E blades was lower than that of NA (P < 0.05). In order to investigate the causes of changes in strain E blades, total RNA in strain E and NA blades were sequenced using the Illumina Hiseq platform. Compared with NA, 1,549 unigenes were selected in strain E including 657 up‐regulated and 892 down‐regulated genes. According to the physiology measurement and differentially expressed genes analysis, cell respiration in strain E might decrease, whereas anabolic‐like photosynthesis and protein biosynthesis might increase compared with NA. This means substance accumulation might be greater than decomposition in strain E. This might explain why strain E blades showed improved growth compared with NA. In addition, several genes related to stress resistance were up‐regulated in strain E indicating that strain E might have a higher stress resistance. The sequencing dataset may be conducive to Pyropia yezoensis molecular breeding research.  相似文献   

7.
Two new polyphenol oxidase (PPO) cDNAs (PPO3 and PPO4 cDNAs, accession numbers GQ354801 and GQ354802, respectively) were obtained by RACE-PCR from Agaricus bisporus. PPO3 cDNA was 1844 bp in length with an open reading frame of 1731 bp, while PPO4 cDNA was 2042 bp with an open reading frame of 1836 bp. PPO3 and PPO4 cDNAs, with 52% identity at the nucleic acid level, encoded a 576-amino acid protein of 66.3 kDa and 611-amino acid protein of 68.3 kDa, respectively. Mature forms of PPO3 and PPO4 were characterized after removing the specific C-terminal region and expressed in Escherichia coli BL21 (DE3) RIPL using pGEX-4T-1 vector. The expressed proteins were probed by the anti-A. bisporus PPO antibody but without PPO activity. This indicated that the recombinant mature PPO3 and mature PPO4 could not form an active center in prokaryotic expression system.  相似文献   

8.
Kinetic measurements of a novel copper-dependent amine oxidase, purified from rat liver mitochondria matrix, were carried out using various substrates in a large pH (5.6–10.2) and ionic strength range (5–200 mM), in order to study the docking of substrates to the enzyme and, as a consequence, to verify the physicochemical characteristics of the active site. Relatively small changes of V max values (approx. 2.5-folds) over the substrates tested, suggest that the rate determining step of the catalysis is only slightly affected by amine chemical structure. In contrast, the strong change of K M and k c/K M values (approx. two orders of magnitude) indicates electrostatic control of the docking process, since the changes of K M and k c/K M values appear due to the presence of positively charged groups in the substrate molecules. These results suggest the presence in the enzyme active site of two negatively charged amino acid residues which seem to interact with positively charged groups of the substrate molecules. Analogies and differences with bovine serum amine oxidase are also described.  相似文献   

9.
Target spot, caused by the fungus Corynespora cassiicola, has become a serious foliar disease in soybean production in the Brazilian Cerrado. Information in the literature regarding the biochemical defence responses of soybean to C. cassiicola infection is rare. Therefore, the objective of this study was to determine the biochemical features associated with soybean resistance to target spot. The activities of chitinases (CHI), β‐1‐3‐glucanases (GLU), phenylalanine ammonia‐lyases (PAL), peroxidases (POX), polyphenol oxidases (PPO) and lipoxygenases (LOX), as well as the concentrations of total soluble phenolics (TSP) and lignin‐thioglycolic acid (LTGA) derivatives, were determined in soybean leaves from both a resistant (FUNDACEP 59) and a susceptible (TMG 132) cultivar. The target spot severity, number of lesions per cm2 of leaflet and area under the disease progress curve were significantly lower for plants from cv. FUNDACEP 59 compared to plants from cv. TMG 132. The GLU, CHI, PAL, POX and PPO activities and the concentration of LTGA derivatives increased significantly, whereas LOX activity decreased significantly on the leaves infected by C. cassiicola. Inoculated plants from cv. FUNDACEP 59 showed a higher PPO activity and concentrations of TSP and LTGA derivatives at 4 and 6 days after inoculation compared to plants from cv. TMG 132. In conclusion, the results of this study demonstrated that the defence‐related enzyme activities increased upon C. cassiicola infection, regardless of the basal level of resistance of the cultivar studied. The increases in PPO activity and concentrations of TSP and LTGA derivatives, but lower LOX activity, at early stages of C. cassiicola infection were highly associated with soybean resistance to target spot.  相似文献   

10.
The differences in the ability of the invading whitefly, Bemisia tabaci (Gennadius) (commonly known as biotype B and hereafter as B) and Trialeurodes vaporariorum (Westwood) (both Hemiptera: Aleyrodidae) to utilize salivary phenol‐oxidizing enzymes – polyphenol oxidase (PPO) and peroxidase (POD) to detoxify plant defensive phenolic compounds were explored. Polyphenol oxidase and POD were found in the saliva of both B and T. vaporariorum. For tomato colonies, the PPO and POD activities in the watery saliva of B were 2.27‐ and 1.34‐fold higher than those of T. vaporariorum. The PPO activities against specific phenolic compounds commonly found in plants were compared. The activities of those from B were significantly greater than those from T. vaporariorum. We also measured PPO activity in both species after they had fed on plants that were undamaged or had been previously damaged with either a plant pathogen [Phytophthora infestans (Mont.) de Bary (Peronosporales)] infection, mechanical damage, B infestation, or exogenous salicylic acid. For B, PPO activities in watery saliva increased 229, 184, 152, and 139% in response to the four treatments, whereas those of T. vaporariorum only increased 133, 119, 113, and 103%, respectively. Biotype B infestation significantly increased the total phenolic content of tomato leaves. Meanwhile, feeding on tomato infestation with B had no significant effect on the survival rate of B, but decreased the survival rate of T. vaporariorum significantly. These results suggest that B has stronger ability utilizing PPO to detoxify high concentrations of phenolics than T. vaporariorum, and this contributes to a significant advantage for B to hold high fitness on plants with induced resistance. Possible roles of salivary PPO in the competition between B and T. vaporariorum are discussed.  相似文献   

11.
PPO (protoporphyrinogen IX oxidase) catalyses the flavin-dependent six-electron oxidation of protogen (protoporphyrinogen IX) to form proto (protoporphyrin IX), a crucial step in haem and chlorophyll biosynthesis. The apparent K(m) value for wild-type tobacco PPO2 (mitochondrial PPO) was 1.17 muM, with a V(max) of 4.27 muM.min(-1).mg(-1) and a catalytic activity k(cat) of 6.0 s(-1). Amino acid residues that appear important for substrate binding in a crystal structure-based model of the substrate docked in the active site were interrogated by site-directed mutagenesis. PPO2 variant F392H did not reveal detectable enzyme activity indicating an important role of Phe(392) in substrate ring A stacking. Mutations of Leu(356), Leu(372) and Arg(98) increased k(cat) values up to 100-fold, indicating that the native residues are not essential for establishing an orientation of the substrate conductive to catalysis. Increased K(m) values of these PPO2 variants from 2- to 100-fold suggest that these residues are involved in, but not essential to, substrate binding via rings B and C. Moreover, one prominent structural constellation of human PPO causing the disease variegate porphyria (N67W/S374D) was successfully transferred into the tobacco PPO2 background. Therefore tobacco PPO2 represents a useful model system for the understanding of the structure-function relationship underlying detrimental human enzyme defects.  相似文献   

12.
Phosphorylation on the activation loop of AGC kinases is typically mediated by PDK1. The precise mechanism for this in‐trans phosphorylation is unknown; however, docking of a hydrophobic (HF) motif in the C‐tail of the substrate kinase onto the N‐lobe of PDK1 is likely an essential step. Using a peptide array of PKA to identify other PDK1‐interacting sites, we discovered a second AGC‐conserved motif in the C‐tail that interacts with PDK1. Since this motif [FD(X)1‐2Y/F] lies in the active site tether region and in PKA contributes to ATP binding, we call it the Adenosine binding (Ade) motif. The Ade motif is conserved as a PDK1‐interacting site in Akt and PRK2, and we predict it will be a PDK1‐interacting site for most AGC kinases. In PKA, the HF motif is only recognized when the turn motif Ser338 is phosphorylated, possibly serving as a phosphorylation “switch” that regulates how the Ade and HF motifs interact with PDK1. These results demonstrate that the extended AGC C‐tail serves as a polyvalent element that trans‐regulates PDK1 for catalysis. Modeling of the PKA C‐tail onto PDK1 structure creates two chimeric sites; the ATP binding pocket, which is completed by the Ade motif, and the C‐helix, which is positioned by the HF motif. Together, they demonstrate substrate‐assisted catalysis involving two kinases that have co‐evolved as symbiotic partners. The highly regulated turn motifs are the most variable part of the AGC C‐tail. Elucidating the highly regulated cis and trans functions of the AGC tail is a significant future challenge.  相似文献   

13.
The aim of this study was to examine how structural properties of different sphingomyelin (SM) analogs affected their substrate properties with sphingomyelinase (SMase) from Bacillus cereus. Using molecular docking and dynamics simulations (for SMase–SM complex), we then attempted to explain the relationship between SM structure and enzyme activity. With both micellar and monolayer substrates, 3O-methylated SM was found not to be degraded by the SMase. 2N-methylated SM was a substrate, but was degraded at about half the rate of its 2NH–SM control. PhytoPSM was readily hydrolyzed by the enzyme. PSM lacking one methyl in the phosphocholine head group was a good substrate, but PSM lacking two or three methyls failed to act as substrates for SMase. Based on literature data, and our docking and MD simulations, we conclude that the 3O-methylated PSM fails to interact with Mg2 + and Glu53 in the active site, thus preventing hydrolysis. Methylation of 2NH was not crucial for binding to the active site, but appeared to interfere with an induced fit activation of the SMase via interaction with Asp156. An OH on carbon 4 in the long-chain base of phytoPSM appeared not to interfere with the 3OH interacting with Mg2 + and Glu53 in the active site, and thus did not interfere with catalysis. Removing two or three methyls from the PSM head group apparently increased the positive charge on the terminal N significantly, which most likely led to ionic interactions with Glu250 and Glu155 adjacent to the active site. This likely interaction could have misaligned the SM substrate and hindered proper catalysis.  相似文献   

14.
Polyphenol oxidase (PPO) has been shown to reduce proteolysis and lipolysis in red clover through deactivation of proteolytic and lipolytic enzymes and/or through formation of protein–phenol–lipid complexes. This experiment investigated the time course of both lipolysis and proteolysis in two red clover lines with different PPO activities with and without addition of a silage inoculant to help understand the action of PPO in the silo, and its potential effects on protein and glycerol-based lipid conservation, and to determine effects of a more rapid pH reduction with inoculation on PPO activity. Four silages were prepared from high or low PPO precision chopped red clover in 60 test-tube-silos, each containing 120 g fresh weight: (a) high PPO red clover without inoculation (H−), (b) low PPO red clover without inoculation (L−), (c) high PPO red clover with inoculation (H+), and (d) low PPO red clover with inoculation (L+). Each treatment had three replicates for each time point of 1, 2, 4, 8 and 90 days. The inoculant used was Lactobacillus plantarum strain L54 applied at a rate of 106 CFU/g fresh weight. Silage pH was reduced (P < 0.001) by inoculation with no effect of PPO. Inoculation had no effect on either lipolysis or free amino acid release, although more (P < 0.01) soluble protein and less (P < 0.01) ammonia-N was in inoculated silages. H silages had a lower level of both proteolysis (release of free amino acids, P < 0.05) and lipolysis (loss of membrane lipid, P < 0.01) than L red clover silages. Results indicate that PPO reduced proteolysis and lipolysis in the silo and that inoculation had no adverse effects on PPO activity.  相似文献   

15.
Protoporphyrinogen IX oxidase (PPO), the last common enzyme of haem and chlorophyll biosynthesis, catalyses the oxidation of protoporphyrinogen IX to protoporphyrin IX. The membrane-embedded flavoprotein is the target of a large class of herbicides. In humans, a defect in PPO is responsible for the dominantly inherited disease variegate porphyria. Here we present the crystal structure of mitochondrial PPO from tobacco complexed with a phenyl-pyrazol inhibitor. PPO forms a loosely associated dimer and folds into an FAD-binding domain of the p-hydroxybenzoate-hydrolase fold and a substrate-binding domain that enclose a narrow active site cavity beneath the FAD and an alpha-helical membrane-binding domain. The active site architecture suggests a specific substrate-binding mode compatible with the unusual six-electron oxidation. The membrane-binding domains can be docked onto the dimeric structure of human ferrochelatase, the next enzyme in haem biosynthesis, embedded in the opposite side of the membrane. This modelled transmembrane complex provides a structural explanation for the uncoupling of haem biosynthesis observed in variegate porphyria patients and in plants after inhibiting PPO.  相似文献   

16.
Elymus cylindricus (2= 6= 42) and E. breviaristatus (2= 6= 42) are distributed in grasslands and deserts of northern and north‐western China. Their genomic constitution and taxonomic status are unclear. Elymus cylindricus was crossed with E. wawawaiensis J.R.Carlson & Barkworth ( StH ), Roegneria grandis Keng ( StY ) and Campeiostachys dahurica (Turcz. ex Griseb.) B.R.Baum, J.L. Y ang & C. Y en var. dahurica ( StYH ). Meiotic pairing in the hybrids E. cylindricus × E. wawawaiensis ( StH ), E. cylindricus × R. grandis ( StY ) and E. cylindricus × C. dahurica var. dahurica ( StYH ) showed on average 10.00, 11.30 and 20.92 bivalents per cell, respectively. Elymus breviaristatus was crossed with C. dahurica var. dahurica ( StYH ) and E. cylindricus. Chromosome pairing in the hybrids of E. breviaristatus × C. dahurica var. dahurica and E. breviaristatus × E. cylindricus showed on average 19.60 and 19.27 bivalents, respectively. Genomic in situ hybridization (GI SH ) revealed the presence of St , Y and H genomes in E. cylindricus and E. breviaristatus. An intergenomic rearrangement was observed in E. cylindricus using GI SH . Meiotic pairing data and GI SH indicated that both E. cylindricus and E. breviaristatus are allohexaploids containing the StYH genomes. Elymus cylindricus and E. breviaristatus should be treated as Campeiostachys dahurica var. cylindrica and Campeiostachys breviaristata, respectively.  相似文献   

17.
2-oxo-4-[(hydroxy)(methyl)phosphinoyl]butyric acid (PPO) is the essential precursor keto acid for the asymmetric biosynthesis of herbicide l -phosphinothricin (l -PPT). Developing a biocatalytic cascade for PPO production with high efficiency and low cost is highly desired. Herein, a d -amino acid aminotransferase from Bacillus sp. YM-1 (Ym DAAT) with high activity (48.95 U/mg) and affinity (Km = 27.49 mM) toward d -PPT was evaluated. To circumvent the inhibition of by-product d -glutamate (d -Glu), an amino acceptor (α-ketoglutarate) regeneration cascade was constructed as a recombinant Escherichia coli (E. coli D), by coupling Ym d -AAT, d -aspartate oxidase from Thermomyces dupontii (TdDDO) and catalase from Geobacillus sp. CHB1. Moreover, the regulation of the ribosome binding site was employed to overcome the limiting step of expression toxic protein TdDDO in E. coli BL21(DE3). The aminotransferase-driven whole-cell biocatalytic cascade (E. coli D) showed superior catalytic efficiency for the synthesis of PPO from d ,l -phosphinothricin (d ,l -PPT). It revealed the production of PPO exhibited high space–time yield (2.59 g L−1 h−1) with complete conversion of d -PPT to PPO at high substrate concentration (600 mM d ,l -PPT) in 1.5 L reaction system. This study first provides the synthesis of PPO from d ,l -PPT employing an aminotransferase-driven biocatalytic cascade.  相似文献   

18.
This study aimed to characterize the activity of ectonucleoside triphosphate diphosphohydrolase (E‐NTPDase; EC 3.6.1.5) in peritoneal cavity cells from BALB/c mice. E‐NTPDase was activated in the presence of both calcium (1.5mM) and magnesium (1.5mM) ions. However, the activity was higher in the presence of Ca2+. A pH of 8.5 and temperature of 37°C were the optimum conditions for catalysis. The apparent Km values were 0.51mM and 0.66mM for the hydrolysis of adenosine triphosphate (ATP) and adenosine diphosphate (ADP), respectively. The Vmax values were 136.4 and 120.8 nmol Pi/min/mg of protein for ATPase and ADPase activity, respectively. Nucleotide hydrolysis was inhibited in the presence of sodium azide (20mM, ATP: P < .05; ADP: P < .001), sodium fluoride (20mM; ATP and ADP: P < .001), and suramin (0.3mM; ATP: P < .01; ADP: P < .05), which is a known profile for NTPDase inhibition. Although all of the diphosphate and triphosphate nucleotides that were tested were hydrolyzed, enzyme activity was increased when adenine nucleotides were used as substrates. Finally, we stress that knowledge of the E‐NTPDase catalytic biochemical properties in mouse peritoneal cavity cells is indispensable for properly determining its activity, as well as to fully understand the immune response profile in both healthy and sick cells.  相似文献   

19.
Efimov I  Cronin CN  McIntire WS 《Biochemistry》2001,40(7):2155-2166
Each flavoprotein subunit (alpha or PchF) of the alpha(2)beta(2) flavocytochrome p-cresol methylhydroxylase (PCMH) from Pseudomonas putida contains FAD covalently attached to Tyr384. PCMH oxidizes p-cresol to 4-hydroxybenzyl alcohol, which is oxidized subsequently by PCMH to 4-hydroxybenzaldehyde. The Y384F mutant form of PchF (apo-PchF[Y384F]) displayed stoichiometric noncovalent FAD binding. PchF[Y384F]FAD associated with the cytochrome subunit (beta or PchC) (producing PCMH[Y384F]), although not as avidly as with wild-type PchF containing covalently bound FAD (PchF(C)). Dramatic increases in the two-electron E(m,7) (NHE) values for FAD were observed when it bound noncovalently to either apo-PchF or apo-PchF[Y384F], and the two-electron E(m,7) value for FAD was increased further by about 75 mV upon covalent binding to PchF, i.e., PchF(C). The E(m,7) values increased by approximately 20 and 45 mV, respectively, when PchF(C) and PchF[Y384F]FAD associated with PchC. The two-electron E(m,7) for covalently bound FAD in PCMH is 84 mV, the highest measured for a flavoprotein. The values for the one-electron redox potentials (E(m,7), NHE) for FAD were measured also for various forms of PchF. Under anaerobiosis, the reduction of PchF[Y384F]FAD by substrates was similar to that observed previously for PchF containing noncovalently bound FAD. Stopped-flow kinetic studies indicated a rapid substrate reduction of the FAD and heme in PCMH[Y384F] which produced PchF[Y384F]FAD(rad) x PchC, the mutant enzyme containing the flavin radical and reduced heme. These experiments also revealed a slow reduction of unassociated PchC(ox) by PchF[Y384F]FAD(rad) x PchC. Steady-state kinetic studies of the reaction of PCMH[Y384F] with p-cresol indicated that the K(m) for this substrate was unchanged relative to that of PCMH, but that the k(cat) was diminished by an order of magnitude. The data indicate that the covalent attachment of FAD to PchF assists catalysis by raising the E(m,7) of the flavin. Contributions to this effect likely result from conformational changes.  相似文献   

20.
Enzyme promiscuity is a prerequisite for fast divergent evolution of biocatalysts. A phosphotriesterase-like lactonase (PLL) from Geobacillus kaustophilus HTA426 (GkaP) exhibits main lactonase and promiscuous phosphotriesterase activities. To understand its catalytic and evolutionary mechanisms, we investigated a “hot spot” in the active site by saturation mutagenesis as well as X-ray crystallographic analyses. We found that position 99 in the active site was involved in substrate discrimination. One mutant, Y99L, exhibited 11-fold improvement over wild-type in reactivity (kcat/Km) toward the phosphotriesterase substrate ethyl-paraoxon, but showed 15-fold decrease toward the lactonase substrate δ-decanolactone, resulting in a 157-fold inversion of the substrate specificity. Structural analysis of Y99L revealed that the mutation causes a ∼6.6 Å outward shift of adjacent loop 7, which may cause increased flexibility of the active site and facilitate accommodation and/or catalysis of organophosphate substrate. This study provides for the PLL family an example of how the evolutionary route from promiscuity to specificity can derive from very few mutations, which promotes alteration in the conformational adjustment of the active site loops, in turn draws the capacity of substrate binding and activity.  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