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.
M Szczesiul  D E Wampler 《Biochemistry》1976,15(10):2236-2244
Six enzymes involved in the conversion of aspartate to threonine have been extracted from Escherichia coli and separated from each other. Two of these enzymes, aspartokinase and homoserine dehydrogenase, have also been partially purified from Rhodopseudomonas spheroides. In an attempt to determine whether small changes in the kinetic properties of individual enzymes are important to the regulation of metabolic flux through a coupled reaction system, the partially purified enzymes were recombined in a variety of ways under reaction conditions designed to resemble the in vivo situation. These conditions include: use of an entire metabolic system rather than a single reaction; high enzyme concentrations at the same relative concentrations as found in the cell; and low, steady-state concentrations of substrates and products. Metabolic flux was followed spectrophotometrically and the concentrations of aspartic semialdehyde, hemoserine, O-phosphohomoserine, and threonine were measured. The results indicate that the threonine concentration is of major importance in regulating metabolic flux by inhibiting aspartokinase, the first reaction in threonine in the pathway. When threonine-insensitive aspartokinases were used, concentrations reached higher levels and the rate of NADPH oxidation remained higher. The fact that neither aspartic semialdehyde nor homoserine accumulated as the threonine concentration increased and the lack of correlation between changes in metabolic flux and ADP/ATP or NADPH/NADP ratios indicate that more subtle forms of metabolic regulation, such as "reverse cascade", secondary feedback sites, or "energy charge", are of little regulatory importance in this isolated, metabolic system. The results also emphasize the need for caution in projecting in vivo control mechanisms from in vitro experiments.  相似文献   

2.
Serine hydroxymethyltransferase: origin of substrate specificity.   总被引:5,自引:0,他引:5  
All forms of serine hydroxymethyltransferase, for which a primary structure is known, have five threonine residues near the active-site lysyl residue (K229) that forms the internal aldimine with pyridoxal phosphate. For Escherichia coli serine hydroxymethyltransferase each of these threonine residues has been changed to an alanine residue. The resulting five mutant enzymes were purified and characterized with respect to kinetic and spectral properties. The mutant enzymes T224A and T227A showed no significant changes in kinetic and spectral properties compared to the wild-type enzyme. The T225A and T230A enzymes exhibited differences in Km and kcat values but exhibited the same spectral properties as the wild-type enzyme. The four threonine residues at positions 224, 225, 227, and 230 do not play a critical role in the mechanism of the enzyme. The T226A enzyme had nearly normal affinity for substrates and coenzymes but had only 3% of the catalytic activity of the wild-type enzyme. The spectrum of the T226A enzyme in the presence of amino acid substrates showed a large absorption maximum at 343 nm with only a small absorption band at 425 nm, unlike the wild-type enzyme whose enzyme-substrate complexes absorb at 425 nm. Rapid reaction studies showed that when amino acid substrates and substrate analogues were added to the T226A enzyme, the internal aldimine absorbing at 422 nm was rapidly converted to a complex absorbing at 343 nm in a second-order process. This was followed by a very slow first-order formation of a complex absorbing at 425 nm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Homoserine dehydrogenase (HSD), which is required for the synthesis of threonine, isoleucine and methionine in fungi, is a potential target for novel antifungal drugs. In order to design effective inhibitors, the kinetic mechanism of Saccharomyces cerevisiae HSD and the stereochemistry of hydride transfer were examined. Product inhibition experiments revealed that yeast HSD follows an ordered Bi Bi kinetic mechanism, where NAD(P)H must bind the enzyme prior to aspartate semialdehyde (ASA) and homoserine is released first followed by NAD(P)+. H-(1,2,4-triazol-3-yl)-D,L-alanine was an uncompetitive inhibitor of HSD with respect to NADPH (K(ii)=3.04+/-0.18 mM) and a noncompetitive inhibitor with respect to ASA (K(is)=1.64+/-0.36 mM, K(ii)=3.84+/-0.46 mM), in agreement with the proposed substrate order. Both kinetic isotope and viscosity experiments provided evidence for a very rapid catalytic step and suggest nicotinamide release to be primarily rate limiting. Incubation of HSD with stereospecifically deuterated NADP[2H] and subsaturating amounts of aspartate semialdehyde revealed that the pro-S NADPH hydride is transferred to the aldehyde. The pH dependence of steady state kinetic parameters indicate that ionizable groups with basic pKs may be involved in substrate binding, consistent with the observation of Lys223 at the enzyme active site in the recently determined 3D structure [B. DeLaBarre, P.R. Thompson, G.D. Wright, A.M. Berghuis, Nat. Struct. Biol. 7 (2000) 238-244]. These findings provide the requisite foundation for future exploitation of fungal HSD in inhibitor design.  相似文献   

4.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

5.
The kinetic scheme of octopine dehydrogenase of Pecten maximus L., a monomeric enzyme obeying a bi-ter sequential mechanism, was completed, essentially in the forward reaction, by steady-state studies over a wide range of substrate concentration at pH 7.0. Deviation from the Michaelis-Menten behavior with respect to NAD+ and other significant kinetic data led us to ascribe for octopine dehydrogenase mechanism the mnemonical enzyme concept. In addition, another regulatory behavior can be envisaged involving the formation of two dead-end complexes enzyme.NADH.D-octopine and enzyme.NAD+.pyruvate.L-arginine.  相似文献   

6.
Various kinetic approaches were carried out to investigate kinetic attributes for the dual coenzyme activities of mitochondrial aldehyde dehydrogenase from rat liver. The enzyme catalyses NAD(+)- and NADP(+)-dependent oxidations of ethanal by an ordered bi-bi mechanism with NAD(P)+ as the first reactant bound and NAD(P)H as the last product released. The two coenzymes presumably interact with the kinetically identical site. NAD+ forms the dynamic binary complex with the enzyme, while the enzyme-NAD(P)H complex formation is associated with conformation change(s). A stopped-flow burst of NAD(P)H formation, followed by a slower steady-state turnover, suggests that either the deacylation or the release of NAD(P)H is rate limiting. Although NADP+ is reduced by a faster burst rate, NAD+ is slightly favored as the coenzyme by virtue of its marginally faster turnover rate.  相似文献   

7.
The threonine-sensitive homoserine dehydrogenase has been isolated and extensively purified from shoots of Zea mays L. var. earliking. This enzyme is shown to be hysteretic under certain conditions. Progress curves of the NAD-dependent reaction catalyzed by the maize enzyme can be characterized by distinct lags prior to achievement of steady state velocities, reflecting transitions from less active species to a more active steady state form of the enzyme. Incubation of the enzyme for 1 min at 25 degrees C prior to initiation of the reaction profoundly influences the properties of the less active enzyme and the nature of the subsequent slow transitions during assay. When the feedback modifier, L-threonine, or KCl is included in the preincubation mixture, the transitions involve biomolecular association reactions. In the absence of either ligand, or in the presence of an appropriate mixture of both, a unimolecular transition occurs during assay. Three unique preincubation states of the enzyme have been identified on the basis of their response to substrates and effectors; whereas, the kinetic and regulatory properties of the steady state form of the enzyme are independent of preincubation conditions. Steady state can thus be achieved by three different transitions. Each transition is retarded by threonine and favored by substrates and potassium, although the effects of these compounds differ quantitatively. Under the conditions tested, monovalent cations have no effect on the steady state velocity of the enzyme. A model describing the relationships among the four unique states of the enzyme which is consistent with the present results and supported by previous observations is proposed.  相似文献   

8.
Matthews  Benjamin F.  Widholm  Jack M. 《Planta》1978,141(3):315-321
Aspartokinase (EC 2.7.2.4), homoserine-dehydrogenase (EC 1.1.1.3) and dihydrodipicolinic-acid-synthase (EC 4.2.1.52) activities were examined in extracts from 1-year-old and 11-year-old cell suspension cultures and whole roots of garden carrot (Daucus carota L.). Aspartokinase activity from suspension cultures was inhibited 85% by 10 mM L-lysine and 15% by 10mM L-threonine. In contrast, aspartokinase activity from whole roots was inhibited 45% by 10 mM lysine and 55% by 10 mM threonine. This difference may be based upon alterations in the ratios of the two forms (lysine-and threonine-sensitive) of aspartokinase, since the activity is consistently inhibited 100% by lysine+threonine. Only one form each of homoserine dehydrogenase and of dihydrodipicolinic acid synthase was found in extracts from cell suspension cultures and whole roots. The regulatory properties of either enzyme were identical from the two sources. In both the direction of homoserine formation and aspartic--semialdehyde formation, homoserine dehydrogenase activities were inhibited by 10mM threonine and 10 mM L-cysteine in the presence of NADH or NADPH. KCl increased homoserine dehydrogenase activity to 185% of control values and increased the inhibitory effect of threonine. Dihydrodipicolinic acid synthase activities from both sources were inhibited over 80% by 0.5 mM lysine. Aspartokinase was less sensitive to inhibition by low concentrations of lysine and threonine than were dihydrodipicolinic acid synthase and homoserine dehydrogenase to inhibition by the respective inhibitors.  相似文献   

9.
Homoserine dehydrogenase from cell suspension cultures of carrot (Daucus carota L.) has been purified to apparent homogeneity by a combination of selective heat denaturation, ion exchange and gel filtration chromatographies, and preparative gel electrophoresis. Carrot homoserine dehydrogenase is composed of subunits of equal molecular weight (85,000 ± 5,000). During purification, the enzyme exists predominantly in two molecular weight forms, 180,000 and 240,000. The enzyme can be reversibly converted from one form to the other, and each has different regulatory properties. When the enzyme is dialyzed in the presence of 5 millimolar threonine, the purified enzyme is converted into its trimeric form (240,000), which is completely inhibited by 5 millimolar threonine and is stimulated 2.6-fold by K+. When the enzyme is dialyzed in the presence of K+ and absence of threonine, the purified enzyme is converted into a dimer (180,000), which is not inhibited by threonine and is only stimulated 1.5-fold by K+. The enzyme also can polymerize under certain conditions to form higher molecular weight aggregates ranging in size up to 720,000, which also are catalytically active. This interconversion of homoserine dehydrogenase conformations may reflect the daily stream of events occurring in vivo. When light stimulates protein synthesis, the threonine pool decreases in the chloroplast, while K+ concentrations increase. The change in threonine and K+ concentrations shift the homoserine dehydrogenase from the threonine-sensitive to the threonine-insensitive conformation resulting in increased production of threonine, which would meet the demands of protein synthesis. The reverse process would occur in the dark.  相似文献   

10.
Two forms of homoserine dehydrogenase exist in the leaves of both barley and pea; one has a large molecular weight and is inhibited by threonine, the other is of smaller molecular weight and insensitive to threonine but inhibited by cysteine. The subcellular localisation of these enzymes has been examined. Both plants have 60–65% of the total homoserine dehydrogenase activity present in the chloroplast and this activity is inhibited by threonine. The low molecular weight, threonine-insensitive form is present in the cytoplasm. Total homoserine dehydrogenase activity from barley leaves showed progressive desensitisation towards threonine with age in a similar manner to that previously described for maize. It was shown that the effect was due to desensitisation of the chloroplast enzyme, and not to an increase in the insensitive cytoplasm enzyme. No corresponding desensitisation to threonine was detected in pea leaves. The different forms of homoserine dehydrogenase could be separated from pea leaves by chromatography on Blue Sepharose; the threonine-sensitive enzyme passed straight through and the threonine insensitive form was bound. A similar separation of the barley leaf isoenzymes was obtained using Matrex Gel Red A affinity columns; in this case however, the threonine-sensitive isoenzyme was bound. In both plants, the threonine insensitive isoenzyme was subject to greater inhibition by cysteine than was the threonine-sensitive isoenzyme.Abbreviation HSDH homoserine dehydrogenase  相似文献   

11.
Born TL  Blanchard JS 《Biochemistry》1999,38(43):14416-14423
The first unique step in bacterial and plant methionine biosynthesis involves the activation of the gamma-hydroxyl of homoserine. In Escherichia coli, this activation is accomplished via a succinylation reaction catalyzed by homoserine transsuccinylase. The activity of this enzyme is closely regulated in vivo and therefore represents a critical control point for cell growth and viability. We have cloned homoserine transsuccinylase from E. coli and present the first detailed enzymatic study of this enzyme. Steady-state kinetic experiments demonstrate that the enzyme utilizes a ping-pong kinetic mechanism in which the succinyl group of succinyl-CoA is initially transferred to an enzyme nucleophile before subsequent transfer to homoserine to form the final product, O-succinylhomoserine. The maximal velocity, V/K(succinyl)(-)(CoA), and V/K(homoserine) all exhibited a bell-shaped pH dependence with apparent pK's of 6.6 and approximately 7.9. The enzyme was inhibited by iodoacetamide in a pH-dependent manner, with an apparent pK of the group being inactivated of 6.4. This suggests the presence of an active site cysteine which forms a succinyl-cysteine intermediate during enzymatic turnover. Solvent kinetic isotope effect studies yielded inverse effects of 0.7 on V and 0.61 on V/K in the reverse reaction only. On the basis of these observations, we propose a detailed chemical mechanism for this important member of the acyltransferase family.  相似文献   

12.
Aspartokinase has been isolated from wheat germ and a preliminary survey made of its properties in a partially purified extract. The enzyme has an absolute requirement for ATP and a divalent metal ion. The phosphate donor can be either ATP or GTP, but other nucleotides are ineffective. Both magnesium and manganese will activate the enzyme, whereas calcium shows a trace amount of activity. The enzyme has a Km of 16.7 mm for aspartate, 1.2 mm for ATP, and 3.3 mm for MgCl(2). Lysine inhibits the reaction at fairly low concentrations, and threonine inhibits at high concentrations. Other amino acids which are derived from aspartate (methionine, homoserine, threonine, and isoleucine) have little effect. When lysine and threonine are added together, they show a concerted inhibition of the reaction. The enzyme is also stabilized against heat inactivation by lysine and threonine together but not by either when added separately. It is suggested that aspartokinase from plants is a regulatory enzyme and exhibits a concerted feedback mechanism.  相似文献   

13.
The irreversible thermal unfolding of the class A beta-lactamase I from Bacillus cereus has been investigated at pH 7.0, using differential scanning calorimetry (DSC) and inactivation kinetic techniques. DSC transitions showed a single peak with a denaturation enthalpy of 646 kJ.mol-1 and were moderately scan rate dependent, suggesting that the process was partially kinetically controlled. The inactivation kinetics at constant temperature showed that the irreversible denaturation of the enzyme occurs as the sum of two exponential terms whose amplitudes are strongly temperature dependent within the transition range so that, at the lowest temperatures within this interval, irreversible inactivation would proceed mainly through the slow phase. The fraction of irreversibly denatured enzyme (D) as a function of temperature for a given scanning rate was calculated by numerical integration of the kinetic equation with temperature, using previously determined kinetic parameters. This D form was the most populated of the unfolded states only at temperatures well above the maximum in the calorimetric transition. Combination of the results of kinetic and DSC experiments has allowed us to separate the contribution of the final D state to the excess enthalpy change from the contribution arising from the reversibly denatured forms of the enzyme (I(i), i = 1,..., n), with the resulting conclusion that the scan rate dependence of the calorimetric traces was the result of two different dynamic effects, viz., the irreversible step and a slow relaxation process during formation of the reversibly denatured intermediate states. Finally, the problems of using results obtained at a single scan rate to validate the two-state kinetic model are commented on.  相似文献   

14.
Born TL  Franklin M  Blanchard JS 《Biochemistry》2000,39(29):8556-8564
The first unique step in bacterial and plant methionine biosynthesis involves the acylation of the gamma-hydroxyl of homoserine. In Haemophilus influenzae, acylation is accomplished via an acetyl-CoA-dependent acetylation catalyzed by homoserine transacetylase. The activity of this enzyme regulates flux of homoserine into multiple biosynthetic pathways and, therefore, represents a critical control point for cell growth and viability. We have cloned homoserine transacetylase from H. influenzae and present the first detailed enzymatic study of this enzyme. Steady-state kinetic experiments demonstrate that the enzyme utilizes a ping-pong kinetic mechanism in which the acetyl group of acetyl-CoA is initially transferred to an enzyme nucleophile before subsequent transfer to homoserine to form the final product, O-acetylhomoserine. The maximal velocity and V/K(homoserine) were independent of pH over the range of values tested, while V/K(acetyl)(-)(CoA) was dependent upon the ionization state of a single group exhibiting a pK value of 8.6, which was required to be protonated. Solvent kinetic isotope effect studies yielded inverse effects of 0.75 on V and 0.74 on V/K(CoA) on the reverse reaction and effects of 1.2 on V and 1.7 on V/K(homoserine) on the forward reaction. Direct evidence for the formation of an acetyl-enzyme intermediate was obtained using rapid-quench labeling studies. On the basis of these observations, we propose a chemical mechanism for this important member of the acyltransferase family and contrast its mechanism with that of homoserine transsuccinylase.  相似文献   

15.
The activity of three enzymes, aspartokinase, homoserine dehydrogenase, and homoserine kinase, has been studied in the industrial strainSaccharomyces cerevisiae IFI256 and in the mutants derived from it that are able to overproduce methionine and/or threonine. Most of the mutants showed alteration of the kinetic properties of the enzymes aspartokinase, which was less inhibited by threonine and increased its affinity for aspartate, and homoserine dehydrogenase and homoserine kinase, which both lost affinity for homoserine. Furthermore, they showed in vitro specific activities for aspartokinase and homoserine kinase that were higher than those of the wild type, resulting in accumulation of aspartate, homoserine, threonine, and/or methionine/S-adenosyl-methionine (Ado-Met). Together with an increase in the specific activity of both aspartokinase and homoserine kinase, there was a considerable and parallel increase in methionine and threonine concentration in the mutants. Those which produced the maximal concentration of these amino acids underwent minimal aspartokinase inhibition by threonine. This supports previous data that identify aspartokinase as the main agent in the regulation of the biosynthetic pathway of these amino acids. The homoserine kinase in the mutants showed inhibition by methionine together with a lack or a reduction of the inhibition by threonine that the wild type undergoes, which finding suggests an important role for this enzyme in methionine and threonine regulation. Finally, homoserine dehydrogenase displayed very similar specific activity in the mutants and the wild type in spite of the changes observed in amino acid concentrations; this points to a minor role for this enzyme in amino acid regulation.  相似文献   

16.
The control of aspartokinase and homoserine dehydrogenase activities was compared in aerobic and fermentative pseudomonads (genera Pseudomonas and Aeromonas), and in coliform bacteria representative of the principal genera of the Enterobacteriaceae. Isofunctional aspartokinases subject to independent end-product control occur in the Enterobacteriaceae and in Aeromonas. In Pseudomonas, there appears to be a single aspartokinase, subject to concerted feedback inhibition by lysine and threonine. Within this genus, the sensitivity of aspartokinase to the single allosteric inhibitors varies considerably: the aspartokinase of the acidovorans group is little affected by the single inhibitors, whereas that of the fluorescent group is severely inhibited by either amino acid at high concentration. In all bacteria examined, homoserine dehydrogenase activity is inhibited by threonine; inhibition is more severe in aerobic pseudomonads than in the other groups. In most of the bacteria examined, either nicotinamide adenine dinucleotide (NAD) or nicotinamide adenine dinucleotide phosphate can serve as a cofactor for this enzyme, though the relative activity with the two pyridine nucleotides varies widely. Aerobic pseudomonads of the acidovorans group contain a homoserine dehydrogenase that is absolutely specific for NAD. The taxonomic implications of these findings are discussed.  相似文献   

17.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase family, catalyzes the first step in the catabolic pathway of the prostaglandins. This enzyme oxidizes the 15-hydroxyl group of prostaglandins to produce 15-keto metabolites which are usually biologically inactive. A relatively conserved threonine residue corresponding to threonine 11 of 15-PGDH is proposed to be involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine the important role of this residue. Threonine 11 was changed to alanine (T11A), cysteine (T11C), serine (T11S) or tyrosine (T11Y) and the mutant proteins were expressed in E. coli. Western-blot analysis showed that the expression levels of mutant proteins were comparable to that of the wild-type enzyme. Mutants T11A, T11C and T11Y were found to be inactive. Mutant T11S still retained substantial activity and the K(m) value for prostaglandin E(2) (PGE(2)) was similar to the wild-type enzyme; however, the K(m) value for NAD(+) was increased over 23-fold. These results suggest that threonine 11 may be involved in the interaction with NAD(+) either directly or indirectly and contributes to the full catalytic activity of 15-PGDH.  相似文献   

18.
Challenging auxotrophs on metabolites that are precursors of a biosynthetic step involving a mutated enzyme has revealed a new class of suppressor mutations which act by derepressing a minor enzyme activity not normally detected in the wild-type strain. These indirect, partial suppressor mutations which allow isoleucine auxotrophs to grow on homoserine or threonine have been analyzed to determine their effect on enzymes involved in the biosynthesis of these amino acids. It has been found that one class of these suppressor mutations (sprA) leads to the derepression of homoserine kinase, homoserine dehydrogenase, and a minor threonine dehydratase that is not sufficiently active to be detected in the wild-type strain. The gene encoding this second threonine dehydratase activity has been found to be located between the structural genes for homoserine kinase and homoserine dehydrogenase. The results of these experiments indicate that plating of auxotrophs on precursors of a biosynthetic step involving mutated enzymes could prove to be a valuable method for the detection of regulatory mutants as well as a possible tool in studying the evolution of biochemical pathways.  相似文献   

19.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH) is the key enzyme in the inactivation pathway of prostaglandins. It is a member of the short-chain dehydrogenase family of enzymes. A relatively conserved threonine residue corresponding to threonine 188 of 15-PGDH is proposed to be involved in the interaction with the carboxamide group of NAD+. Site-directed mutagenesis was used to examine the important role of this residue. Threonine 188 was changed to alanine (T188A), serine (T188S) or tyrosine (T188Y) and the mutant proteins were expressed in E. coli. Western blot analysis showed that the expression levels of mutant proteins were similar to that of the wild type protein. Mutants T188A and T188Y were found to be inactive. Mutant T188S still retained substantial activity and the Km value for PGE2 was similar to the wild enzyme; however, the Km value for NAD+ was increased over 100 fold. These results suggest that threonine 188 is critical for interaction with NAD+ and contributes to the full catalytic activity of 15-PGDH.  相似文献   

20.
The inactivations of P450 2B4 and the T302A mutant of 2B4 by tert-butyl acetylene (tBA) and the inactivation of 2B4 T302A by tert-butyl 1-methyl-2-propynyl ether (tBMP) have been investigated. tBA and tBMP inactivated both enzymes in a mechanism-based manner with the losses in enzymatic activity corresponding closely to losses in P450 heme. HPLC and ESI-LC-MS analysis detected two different tBA- or tBMP-modified heme products with masses of 661 and 705 Da, respectively. Interestingly, the inactivations of the P450s 2B4 by tBA and tBMP were partially reversible by dialysis, and the tBA- or tBMP-modified heme products could only be observed with ESI-LC-MS/MS when the inactivated samples were acidified prior to analysis, indicating a requirement for protons in the formation of stable heme adducts in both the wild-type and mutant 2B4 enzymes. Results of studies using artificial oxidants to support enzyme inactivation suggest that the oxenoid-iron activated oxygen species is preferentially utilized during the inactivation of the P450s 2B4 by tBA. These results argue against the use of a peroxo-iron species by P450 2B4 T302A. Molecular dynamics studies of wild-type P450 2B4 reveal that contiguous hydrogen bond networks, including structural waters, link a conserved glutamate (E301) to the distal oxygen of the peroxo-heme species via threonine 302. Interestingly, models of 2B4 T302A reveal that a compensatory, ordered hydrogen bond network forms despite the removal of T302. These results indicate that while T302 may play a role in proton delivery in the formation of the oxenoid-iron complex and in the stabilization of acetylene heme adducts in 2B4, it is not essential for proton delivery given the presence of E301 in the 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