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 31 毫秒
1.
One basic peroxidase isoenzyme, with a pI of 8.8, is present in the intercellular washing fluid in the aerial part of 6-day-old Lupinus albus hypocotyl seedlings. This isoenzyme, called LuP-B2, is the principal soluble component secreted into the apoplastic space and it is a constitutive enzyme along the whole length of etiolated hypocotyl. The enzymatic inactivation process which this apoplastic peroxidase undergoes is described for the first time. The kinetic constants which describe its inactivation by H(2)O(2) in the absence of reductant substrates are determined. LuP-B2 is inactivated in situ and in vitro in a time- and concentration-dependent manner. H(2)O(2) acts as a suicide substrate according to a model previously proposed by us. The constant values calculated are similar to those calculated for the basic isoenzyme of horseradish roots, HRP-C. LuP-B2 presents a k(inact) value of 7.5 x 10(-3) s(-1) and a k(cat) of 6.7 s(-1). This isoenzyme makes 889 catalytic cycles for each inactivation event. The similarity in behavior and the constant values, together with other situations (both are excreted, soluble and constitutive isoenzymes) suggest that the inactivation process could play an important role in plant development and stress situations.  相似文献   

2.
Three genes encoding heme hydroperoxidases (katA, katB, and katC) have been identified in the soil bacterium Sinorhizobium meliloti. The recombinant proteins were overexpressed in Escherichia coli and purified in order to achieve a spectral and kinetic characterization. The three proteins contain heme b with high-spin Fe(III). KatB is an acidic bifunctional homodimeric catalase-peroxidase exhibiting both catalase (k(cat) = 2400 s(-1)) and peroxidase activity and having a high affinity for hydrogen peroxide (apparent K(M) = 1.6 mM). KatA and KatC are acidic monofunctional homotetrameric catalases. Although different in size (KatA is a small subunit catalase while KatC is a large subunit catalase) both enzymes exhibit the same heme type and a similar affinity for H(2)O(2) (apparent K(M) values of 160 and 150 mM). However, the turnover rate of KatA (k(cat) = 279000 s(-1)) exceeds that of KatC (k(cat) = 3100 s(-1)) significantly. The kinetic parameters are in good agreement with the physiological role of these heme proteins. KatB is the housekeeping hydroperoxidase exhibiting the highest affinity for hydrogen peroxide, while KatA has the lowest H(2)O(2) affinity but the highest k(cat)/K(M) value (1.75 x 10(6) M(-1) s(-1)), in agreement with the hydrogen peroxide inducibility of the encoding gene. Moreover, the lower catalytic efficiency of KatC (2.1 x 10(4) M(-1) s(-1)) appears to be enough for growing in the stationary phase and/or under heat or salt stress (conditions that are known to favor katC expression).  相似文献   

3.
Cytidine 5'-triphosphate synthase (CTPS) catalyzes the ATP-dependent formation of CTP from UTP using either NH3 or L-glutamine as the source of nitrogen. To identify the location of the ATP-binding site within the primary structure of E. coli CTPS, we used the affinity label 2',3'-dialdehyde adenosine 5'-triphosphate (oATP). oATP irreversibly inactivated CTPS in a first-order, time-dependent manner while ATP protected the enzyme from inactivation. In the presence of 10 mM UTP, the values of k(inact) and K(I) were 0.054 +/- 0.001 min(-1) and 3.36 +/- 0.02 mM, respectively. CTPS was labeled using (2,8-3H)oATP and subsequently subjected to trypsin-catalyzed proteolysis. The tryptic peptides were separated using reversed-phase HPLC, and two peptides were identified using N-terminal sequencing (S(492)GDDQLVEIIEVPNH(506) and Y(298)IELPDAY(K(306)) in a 5:1 ratio). The latter suggested that Lys 306 had been modified by oATP. Replacement of Lys 306 by alanine reduced the rate of oATP-dependent inactivation (k(inact) = 0.0058 +/- 0.0005 min(-1), K(I) = 3.7 +/- 1.3 mM) and reduced the apparent affinity of CTPS for both ATP and UTP by approximately 2-fold. The efficiency of K306A-catalyzed glutamine-dependent CTP formation was also reduced 2-fold while near wild-type activity was observed when NH3 was the substrate. These findings suggest that Lys 306 is not essential for ATP binding, but does play a role in bringing about the conformational changes that mediate interactions between the ATP and UTP sites, and between the ATP-binding site and the glutamine amide transfer domain. Replacement of the nearby, fully conserved Lys 297 by alanine did not affect NH3-dependent CTP formation, relative to wild-type CTPS, but reduced k(cat) for the glutaminase activity 78-fold. Our findings suggest that the conformational change associated with binding ATP may be transmitted through the L10-alpha11 structural unit (residues 297-312) and thereby mediate effects on the glutaminase activity of CTPS.  相似文献   

4.
N-type inactivation of rat Kv1.4 channels with one, two, or four inactivation balls was investigated using homogeneous populations of channels expressed in Xenopus oocytes. Tandem dimeric and tetrameric constructs of Kv1.4 were made. Channels encoded by tandem cDNAs Kv1. 4-Kv1.4Delta1-145 and Kv1.4-[Kv1.4Delta1-145](3) have two or only one tethered inactivation ball, respectively, whereas Kv1.4 itself encodes channels having four inactivation balls. The time constants for inactivation of macroscopic currents were increased significantly as the number of inactivation balls was decreased, whereas the time constants for recovery from inactivation were not modified. The ratios of the rate constants of inactivation (k(inact)) of Kv1.4-Kv1.4Delta1-145 and Kv1.4-[Kv1.4Delta1-145](3) channels to that of the Kv1.4 channel were 0.65 and 0.4, respectively, whereas the ratios of the rate constant of recovery (k(rec)) of these channels to that of Kv1.4 were almost unity. The rate constants k(inact) for channels having two and four inactivation balls are smaller than those that would be expected if inactivation balls on each channel are independent, suggesting some interaction occurs between inactivation balls. Furthermore, noninactivating current became apparent as the number of inactivation balls on a channel was decreased.  相似文献   

5.
26,27-dehydrocycloartenol (26,27-DHC) was shown to be a substrate for the soybean sterol methyltransferase (SMT) as well as a mechanism-based inhibitor of enzyme action. The K(m) and k(cat) for 26,27-DHC was 10 microM and 0.018 min(-1), respectively. SMT catalyzed 26,27-DHC to two products tentatively identified as 26-homocholesta-9,19-cyclo-23(24)E,26(26')-dienol and 26-homocholesta-9,19-cyclo-26(26')-en-3beta,24beta-diol by GC-MS. Inhibitor treatment was concentration- and time-dependent (pseudo-first-order kinetics). A replot of the half-lives for inactivation versus the inverse of the inactivator concentrations gave an apparent K(i) of 42 microM and a maximum rate of inactivation of 0.29 min(-1). A partition ratio (k(cat)/k(inact)) was calculated to be 0.06.  相似文献   

6.
Dai Y  Hedstrom L  Abeles RH 《Biochemistry》2000,39(21):6498-6502
We have synthesized (acyloxy)methyl ketone inactivators of papain, cathepsin B, and interleukin-1beta conversion enzyme (ICE) that interact with both the S and S' subsites. The value of k(inact)/K(i) for these inactivators is strongly dependent on the leaving group. For example, Z-Phe-Gly-CH(2)-X is a poor inactivator of papain when X is OCOCH(3) (k(inact)/K(i) = 2.5 M(-)(1) s(-)(1)) but becomes a potent inactivator when X is OCO-L-Leu-Z (k(inact)/K(i) = 11 000 M(-)(1) s(-)(1)). Since these leaving groups have similar chemical reactivities, the difference in potency must be attributed to interactions with the S' sites. The potency of the leaving group correlates with the P' specificity of papain. Similar results are also observed for the inactivation of cathepsin B by these compounds. A series of inactivators with the general structure Fmoc-L-Asp-CH(2)-X were designed to inactivate ICE. No inhibition was observed when X was OCOCH(3). In contrast, ICE is inactivated when X is OCO-D-Pro-Z (k(inact)/K(i) = 131 M(-)(1) s(-)(1)). These results demonstrate that S'-P' interactions can be utilized to increase the efficacy and selectivity of (acyloxy)methyl ketone inactivators.  相似文献   

7.
Prostaglandin endoperoxide H synthase (PGHS) catalyzes the oxidation of arachidonate to prostaglandin H(2). We have previously synthesized and chemically characterized nitroarachidonic acid (AANO(2)), a novel anti-inflammatory signaling mediator. Herein, the interaction of AANO(2) with PGHS was analyzed. AANO(2) inhibited oxygenase activity of PGHS-1 but not PGHS-2. AANO(2) exhibited time- and concentration-dependent inhibition of peroxidase activity in both PGHS-1 and -2. The plot of k(obs) versus AANO(2) concentrations showed a hyperbolic function with k(inact) = 0.045 s(-1) and K(i)(*app) = 0.019 μM for PGHS-1 and k(inact) = 0.057 s(-1) and K(i)(*app) = 0.020 μM for PGHS-2. Kinetic analysis suggests that inactivation of PGHS by AANO(2) involves two sequential steps: an initial reversible binding event (described by K(i)) followed by a practically irreversible event (K(i)(*app)) leading to an inactivated enzyme. Inactivation was associated with irreversible disruption of heme binding to the protein. The inhibitory effects of AANO(2) were selective because other nitro-fatty acids tested, such as nitrooleic acid and nitrolinoleic acid, were unable to inhibit enzyme activity. In activated human platelets, AANO(2) significantly decreased PGHS-1-dependent thromboxane B(2) formation in parallel with a decrease in platelet aggregation, thus confirming the biological relevance of this novel inhibitory pathway.  相似文献   

8.
The apparent catalytic constant (k(cat)) of artichoke (Cynara scolymus L.) peroxidase (AKPC) with 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) (ABTS) increased 130-fold in the presence of calcium ions (Ca2+) but the affinity (K(m)) of the enzyme for ABTS was 500 times lower than for Ca2+-free AKPC. AKPC is known to exhibit an equilibrium between 6-aquo hexa-coordinate and penta-coordinate forms of the haem iron that is modulated by Ca2+ and affects compound I formation. Measurements of the Ca2+ dissociation constant (K(D)) were complicated by the water-association/dissociation equilibrium yielding a global value more than 1000 times too high. The value for the Ca2+ binding step alone has now been determined to be K(D) approximately 10 nM. AKPC-Ca2+ was more resistant to inactivation by hydrogen peroxide (H(2)O(2)) and exhibited increased catalase activity. An analysis of the complex H(2)O(2) concentration dependent kinetics of Ca2+-free AKPC is presented.  相似文献   

9.
Chan WK  Delucchi AB 《Life sciences》2000,67(25):3103-3112
Resveratrol, a phytoalexin found in red wine, has been shown to possess antioxidant and antimutagenic properties. Incubation of resveratrol with Sf9 insect microsomes containing baculovirus-derived human cytochrome P450 3A4 (CYP3A4) and NADPH-cytochrome P450 reductase showed that resveratrol inactivated CYP3A4 in a time- and NADPH-dependent manner. Resveratrol, erythromycin and troleandomycin inactivated CYP3A4 at a similar rate (as reflected by k(inact)) whereas the binding affinity to CYP3A4 (as reflected by K(I)) was in the order of: troleandomycin > erythromycin > resveratrol. (K(I) and k(inact) for CYP3A4 inactivation by resveratrol, erythromycin and troleandomycin are 20 microM and 0.20 min(-1), 5.3 microM and 0.12 min(-1) and 0.18 microM and 0.15 min(-1), respectively.) Fractionation studies of red wine showed that fractions that did not contain resveratrol inactivated CYP3A4 significantly. In addition, the resveratrol content in red wine used in the study was too low to account for the degree of CYP3A4 inactivation observed after red wine treatment. Inactivation studies using a variety of red wine types showed that the CYP3A4 inactivation did not correlate to their resveratrol content. In summary, data here showed that resveratrol is an effective mechanism-based inactivator of CYP3A4; however, it is not one of the main red wine constituents that are responsible for CYP3A4 inactivation by red wine. Nevertheless, inactivation of CYP3A4 by resveratrol may cause clinically relevant drug interactions with CYP3A4 substrates.  相似文献   

10.
Polymers and copolymers of horseradish root peroxidase (HRP) and Penicillium funiculosum 46.1 glucose oxidase (GO) have been synthesized and their catalytic properties have been characterized (free and immobilized forms of each enzyme were studied). The cooxidation reaction of phenol and 4-aminoantipyrin (4-AAP), performed in an aqueous medium in the presence of equimolar amounts of GO and HRP, was characterized by effective K(M) and k(cat) of 0.58 mM and 20.9 s(-1) (for phenol), and 14.6 mM and 18.4 s(-1) (glucose), respectively. The catalytic efficiency of polymerization products (PPs) of GO (GO-PPs) depended on the extent of their aggregation. The combinations GO + HRP-PP and HRP + GO-PP, as well as the copolymer HRP*-GO-PP, proved promising as reagents for enzyme-based analytical systems. When adsorbed on aluminum hydroxide gels, GO-PPs exhibited higher catalytic activity than the non-polymeric enzyme. Maximum retention of GO-PP activity on the inorganic carrier was observed in the case of GO-PP copolymers with an activated HRP. Polymerization of HRP in the presence of a zinc hydroxide gel, paralleled by HRP-PP immobilization onto the gel, increased both the activity of the enzyme and its operational stability.  相似文献   

11.
Ca(v)2.1 mediates voltage-gated Ca2+ entry into neurons and the release of neurotransmitters at synapses of the central nervous system. An inactivation process that is modulated by the auxiliary beta-subunits regulates Ca2+ entry through Ca(v)2.1. However, the molecular mechanism of this alpha1-beta-subunit interaction remains unknown. Herein we report the identification of new determinants within segment IVS6 of the alpha(1)2.1-subunit that markedly influence channel inactivation. Systematic substitution of residues within IVS6 with amino acids of different size, charge, and polarity resulted in mutant channels with rates of fast inactivation (k(inact)) ranging from a 1.5-fold slowing in V1818I (k(inact) = 0.98 +/- 0.09 s(-1) compared with wild type alpha(1)2.1/alpha2-delta/beta1a k(inact) = 1.35 +/- 0.25 s(-1) to a 75-fold acceleration in mutant M1811Q (k(inact) = 102 +/- 3 s(-1). Coexpression of mutant alpha(1)2.1-subunits with beta(2a) resulted in two different phenotypes of current inactivation: 1) a pronounced reduction in the rate of channel inactivation or 2) an attenuation of a slow component in I(Ba) inactivation. Simulations revealed that these two distinct inactivation phenotypes arise from a beta2a-subunit-induced destabilization of the fast-inactivated state. The IVS6- and beta2a-subunit-mediated effects on Ca(v)2.1 inactivation are likely to occur via independent mechanisms.  相似文献   

12.
Peroxidases (EC 1.11.1.7; donor: hydrogen peroxide oxidoreductase) are part of a large group of enzymes. In this study, peroxidase, a primer antioxidant enzyme, was purified with 19.3 fold and 0.2% efficiency from cauliflower (Brassica oleracea L.) by ammonium sulphate precipitation, dialysis, CM-Sephadex ion-exchange chromatography and Sephadex G-25 purification steps. The substrate specificity of peroxidase was investigated using 2,2'-azino-bis(3-ethylbenz-thiazoline-6-sulphonic acid) (ABTS), 2-methoxyphenol (guaiacol), 1,2-dihydroxybenzene (catechol), 1,2,3-trihyidroxybenzene (pyrogallol) and 4-methylcatechol. Also, optimum pH, optimum temperature, optimum ionic strength, stable pH, stable temperature, thermal inactivation conditions were determined for guaiacol/H(2)O(2), pyrogallol/H(2)O(2), ABTS/H(2)O(2), catechol/H(2)O(2) and 4-methyl catechol/H(2)O(2) substrate patterns. The molecular weight (M(w)) of this enzyme was found to be 44 kDa by gel filtration chromatography method. Native polyacrylamide gel electrophoresis (PAGE) was performed for isoenzyme determination and a single band was observed. K(m) and V(max) values were calculated from Lineweaver-Burk graph for each substrate patterns.  相似文献   

13.
Streit BR  DuBois JL 《Biochemistry》2008,47(19):5271-5280
Chlorite dismutase carries out the heme-catalyzed decomposition of ClO2- to Cl- and O2, an unusual transformation with biotechnological and bioremediative applications. The enzyme has been successfully overexpressed for the first time in highly functional form in Escherichia coli and its steady state kinetics studied. The purified enzyme is abundant (55 mg/L cell culture), highly active (approximately 4.7 x 10(3) micromol of ClO2- min(-1) mg(-1) subunit) and nearly stoichiometric in heme; further, it shares spectroscopic and physicochemical features with chlorite dismutases previously isolated from three organisms. A careful study of the enzyme's steady state kinetics has been carried out. ClO2- consumption and O2 release rates were measured, yielding comparable values of kcat (4.5 x 10(5) min(-1)), K(m) (approximately 215 microM), and kcat/Km (3.5 x 10(7) M(-1) s(-1) via either method (4 degrees C, pH 6.8; all values referenced per heme-containing subunit). ClO2-:O2 stoichiometry exhibited a 1:1 relationship under all conditions measured. Though the value of kcat/Km indicates near diffusion control of the reaction, viscosogens had no effect on k(cat)/K(m) or V(max). The product O2 did not inhibit the reaction at saturating [O2], but Cl- is a mixed inhibitor with relatively high values of KI (225 mM for enzyme and 95.6 mM for the enzyme-substrate complex), indicating a relatively low affinity of the heme iron for halogen ions. Chlorite irreversibly inactivates the enzyme after approximately 1.7 x 10(4) turnovers (per heme) and with a half-life of 0.39 min, resulting in bleaching of the heme chromophore. The inactivation K(I) (K(inact)) of 166 microM is similar in magnitude to Km, consistent with a common Michaelis complex on the pathway to both reaction and inactivation. The one-electron peroxidase substrate guaiacol offers incomplete protection of the enzyme from inactivation. Mechanisms in keeping with the available data and the properties of other well-described heme enzymes are proposed.  相似文献   

14.
The effect of a water-miscible ionic liquid, 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4]), on the horseradish peroxidase (HRP)-catalyzed oxidation of 2-methoxyphenol (guaiacol) with hydrogen peroxide (H2O2) was investigated. HRP maintains its high activity in the aqueous mixtures containing various concentrations of the ionic liquid and even in 90% (v/v) ionic liquid. In order to minimize the effect of solution viscosity on the kinetic constants of HRP catalysis, the enzymatic reactions in the subsequent kinetic study were performed in water-ionic liquid mixtures containing 25% (v/v) ionic liquid at maximum. As the concentration of [BMIM][BF4] increased for the oxidation of guaiacol by HRP, the K(m) value increased with a slight decrease in the k(cat) value: The K(m) value increased from 2.8 mM in 100% (v/v) water to 22.5 mM in 25% (v/v) ionic liquid, indicating that ionic liquid significantly weakens the binding affinity of guaiacol to HRP.  相似文献   

15.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

16.
The 10-cyclopropylidene analog of farnesyl diphosphate was shown to be a mechanism-based inhibitor of trichodiene synthase with an inactivation rate (k(inact)) of 0.010 +/- 0.0003 min(-1) and an apparent Ki of 663 +/- 75 nM. The presence of three anomalous sesquiterpene products detected in incubation mixtures indicate that the compound also serves as a substrate of the enzyme.  相似文献   

17.
Wang H  Vath GM  Gleason KJ  Hanna PE  Wagner CR 《Biochemistry》2004,43(25):8234-8246
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from acetyl coenzyme A (AcCoA) to arylamines, hydrazines, and their N-hydroxylated arylamine metabolites. The recently determined three-dimensional structures of prokaryotic NATs have revealed a cysteine protease-like Cys-His-Asp catalytic triad, which resides in a deep and hydrophobic pocket. This catalytic triad is strictly conserved across all known NATs, including hamster NAT2 (Cys-68, His-107, and Asp-122). Treatment of NAT2 with either iodoacetamide (IAM) or bromoacetamide (BAM) at neutral pH rapidly inactivated the enzyme with second-order rate constants of 802.7 +/- 4.0 and 426.9 +/- 21.0 M(-1) s(-1), respectively. MALDI-TOF and ESI mass spectral analysis established that Cys-68 is the only site of alkylation by IAM. Unlike the case for cysteine proteases, no significant inactivation was observed with either iodoacetic acid (IAA) or bromoacetic acid (BAA). Pre-steady state and steady state kinetic analysis with p-nitrophenyl acetate (PNPA) and NAT2 revealed a single-exponential curve for the acetylation step with a second-order rate constant of (1.4 +/- 0.05) x 10(5) M(-1) s(-1), followed by a slow linear rate of (7.85 +/- 0.65) x 10(-3) s(-1) for the deacetylation step. Studies of the pH dependence of the rate of inactivation with IAM and the rate of acetylation with PNPA revealed similar pK(a)(1) values of 5.23 +/- 0.09 and 5.16 +/- 0.04, respectively, and pK(a)(2) values of 6.95 +/- 0.27 and 6.79 +/- 0.25, respectively. Both rates reached their maximum values at pH 6.4 and decreased by only 30% at pH 9.0. Kinetic studies in the presence of D(2)O revealed a large inverse solvent isotope effect on both inactivation and acetylation of NAT2 [k(H)(inact)/k(D)(inact) = 0.65 +/- 0.02 and (k(2)/K(m)(acetyl))(H)/(k(2)/K(m)(acetyl))(D) = 0.60 +/- 0.03], which were found to be identical to the fractionation factors (Phi) derived from proton inventory studies of the rate of acetylation at pL 6.4 and 8.0. Substitution of the catalytic triad Asp-122 with either alanine or asparagine resulted in the complete loss of protein structural integrity and catalytic activity. From these results, it can be concluded that the catalytic mechanism of NAT2 depends on the formation of a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)). However, in contrast to the case with cysteine proteases, a pH-dependent protein conformational change is likely responsible for the second pK(a), and not deprotonation of the thiolate-imidazolium ion. In addition, substitutions of the triad aspartate are not tolerated. The enzyme appears, therefore, to be engineered to rapidly form a stable acetylated species poised to react with an arylamine substrate.  相似文献   

18.
The inactivation of horseradish peroxidase A2 (HRP-A2) with H2O2 as the sole substrate has been studied. In incubation experiments it was found that the fall in HRP-A2 activity was non-linearly dependent on H2O2 concentrations and that a maximum level of inactivation of approximately 80% (i.e. approximately 20% residual activity) was obtained with 2,000 or more equivalents of H2O2. Further inactivation was only induced at much higher H2O2 concentrations. Spectral changes during incubations of up to 5 days showed the presence of a compound III-like species whose abundance was correlated to the level of resistance observed. Inactivation was pH dependent, the enzyme being much more sensitive under acid conditions. A partition ratio (r1 approximately equals 1,140 at pH 6.5) between inactivation and catalysis was calculated from the data. The kinetics of inactivation followed single exponential time curves and were H2O2 concentration dependent. The apparent maximum rate constant of inactivation was lambdamax=3.56+/-0.07x10(-4)s(-1) and the H2O2 concentration required to give lambdamax/2 was K2=9.94+/-0.52 mM. The relationship lambdamax相似文献   

19.
The effects of high-pressure treatment on the reaction rates of horseradish peroxidase (HRP) with guaethol or guaiacol as a hydrogen donor were evaluated from direct transmission measurements in a high-pressure optical cell at 435 nm. Peroxidases are known to be very barostable and insensitive to heat. With guaethol the reaction velocity was independent of pressure up to 500 MPa, but with guaiacol the cytochrome c oxidase underwent a mechanism-based irreversible inhibition of catalytic activity when subjected to pressure; in the resting states (fully oxidized or reduced), it was insensitive to pressure. The enzyme inactivation took place with an inactivation rate constant of 5.15 x 10(-1) min(-1) at 500 MPa, 25 degrees C and pH 7. The degree of inactivation was correlated to the concentration of guaiacol. This is the first report on a mechanism-based pressure inactivation of HRP triggered at moderate pressure and temperature and mediated by the hydrogen donor.  相似文献   

20.
Mann S  Ploux O 《The FEBS journal》2006,273(20):4778-4789
Diaminopelargonic acid aminotransferase (DAPA AT), which is involved in biotin biosynthesis, catalyzes the transamination of 8-amino-7-oxononanoic acid (KAPA) using S-adenosyl-l-methionine (AdoMet) as amino donor. Mycobacterium tuberculosis DAPA AT, a potential therapeutic target, has been overproduced in Escherichia coli and purified to homogeneity using a single efficient step on a nickel-affinity column. The enzyme shows an electronic absorption spectrum typical of pyridoxal 5'-phosphate-dependent enzymes and behaves as a homotetramer in solution. The pH profile of the activity at saturation shows a single ionization group with a pK(a) of 8.0, which was attributed to the active-site lysine residue. The enzyme shows a Ping Pong Bi Bi kinetic mechanism with strong substrate inhibition with the following parameters: K(mAdoMet) = 0.78 +/- 0.20 mm, K(mKAPA) = 3.8 +/- 1.0 microm, k(cat) = 1.0 +/- 0.2 min(-1), K(iKAPA) = 14 +/- 2 microm. Amiclenomycin and a new analogue, 4-(4c-aminocyclohexa-2,5-dien-1r-yl)propanol (referred to as compound 1), were shown to be suicide substrates of this enzyme, with the following inactivation parameters: K(i) = 12 +/- 2 microm, k(inact) = 0.35 +/- 0.05 min(-1), and K(i) = 20 +/- 2 microm, k(inact) = 0.56 +/- 0.05 min(-1), for amiclenomycin and compound 1, respectively. The inactivation was irreversible, and the partition ratios were 1.0 and 1.1 for amiclenomycin and compound 1, respectively, which make these inactivators particularly efficient. compound 1 (100 microg.mL(-1)) completely inhibited the growth of an E. coli C268bioA mutant strain transformed with a plasmid expressing the M. tuberculosis bioA gene, coding for DAPA AT. Reversal of the antibiotic effect was observed on the addition of biotin or DAPA. Thus, compound 1 specifically targets DAPA AT in vivo.  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