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 234 毫秒
1.
Lignin and Mn peroxidase-catalyzed oxidation of phenolic lignin oligomers   总被引:2,自引:0,他引:2  
Banci L  Ciofi-Baffoni S  Tien M 《Biochemistry》1999,38(10):3205-3210
The oxidation of phenolic oligomers by lignin and manganese peroxidases was studied by transient-state kinetic methods. The reactivity of peroxidase intermediates compound I and compound II was studied with the phenol guaiacol along with a beta-O-4 phenolic dimer, trimer, and tetramer. Compound I of both peroxidases is much more reactive than compound II. The rate constants for these substrates with Mn peroxidase compound I range from 1.0 x 10(5) M-1 s-1 for guaiacol to 1.1 x 10(3) M-1 s-1 for the tetramer. Reactivity is much higher with lignin peroxidase compound I with rate constants ranging from 1.2 x 10(6) M-1s-1 for guaiacol to 3.6 x 10(5) M-1 s-1 for the tetramer. Rate constants with compound II are much lower with Mn peroxidase exhibiting very little reactivity. The rate constants dramatically decreased with both peroxidases as the size of the substrate increased. The extent of the decrease was much more dramatic with Mn peroxidase, leading us to conclude that, despite its ability to oxidize phenols, Mn2+ is the only physiologically significant substrate. The rate decrease associated with increasing substrate size was more gradual with lignin peroxidase. These data indicate that whereas Mn peroxidase cannot efficiently directly oxidize the lignin polymer, lignin peroxidase is well suited for direct oxidation of polymeric lignin.  相似文献   

2.
Compared to other monomeric heme proteins and the heme peroxidases, the Glycera dibranchiata monomer hemoglobin components III and IV exhibit very slow cyanide binding kinetics. This is agreement with the previously reported behavior of component II. Similar to component II, components III and IV have been studied under pseudo-first-order conditions at pH 6.0, 7.0, 8.0, and 9.0 by using a 100-250-fold excess of potassium cyanide at each pH. At 20 degrees C with micromolar protein concentrations, kobs for component III varies between 7.08 x 10(-5) s-1 at pH 6.0 and 100-fold cyanide excess and 1.06 x 10(-2) s-1 at pH 9.0 and 250-fold cyanide excess. For component IV, the values are 2.03 x 10(-4) s-1 for 100-fold cyanide excess at pH 6.0 and 4.13 x 10(-2) s-1 for 250-fold cyanide excess at pH 9.0. In comparison to other heme proteins, our analysis shows that the bimolecular rate constant (klapp) is small. For example, at pH 7.0, it is 3.02 x 10(-1) M-1 s-1 for component III and 1.82 M-1 s-1 for component IV, compared to 400 M-1 s-1 for sperm whale metmyoglobin, 692 M-1 s-1 for soybean metleghemoglobin a, 111 M-1 s-1 for guinea pig methemoglobin, and 1.1 x 10(5) M-1 s-1 for cytochrome c peroxidase. Our results also show that the dissociation rates (k-lapp) are extremely slow and no larger than 10(-6) s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
1,2-Dehydro-N-acetyldopamine (dehydro-NADA) is an important catecholamine derivative involved in the cross-linking of insect cuticular components during sclerotization. Since sclerotization is a vital process for the survival of insects, and is closely related to melanogenesis, it is of interest to unravel the chemical mechanisms participating in this process. The present paper reports on the mechanism by which dehydro-NADA is oxidatively activated to form reactive intermediate(s) as revealed by pulse radiolysis, electron spin resonance spectroscopy, high performance liquid chromatography, and ultraviolet-visible spectroscopic analysis. Pulse radiolytic one-electron oxidation of dehydro-NADA by N3. (k = 5.3 x 10(9) M-1 s-1) or Br2.- (k = 7.5 x 10(8) M-1 s-1) at pH6 resulted in the rapid generation of the corresponding semiquinone radical, lambda max 400 nm, epsilon = 20,700 M-1 cm-1. This semiquinone decayed to form a second transient intermediate, lambda max 485 nm, epsilon = 8000 M-1 cm-1, via a second order disproportionation process, k = 6.2 x 10(8) M-1 s-1. At pH 6 in the presence of azide, the first order decay of this second intermediate occurred over milliseconds; the rate decreases at higher pH. At pH 6 in the presence of bromide, the intermediate decayed much more slowly over seconds, k = 0.15 s-1. Under such conditions, the dependence of the first order decay constant upon parent dehydro-NADA concentration led to a second order rate constant of 8.5 x 10(2) M-1 s-1 for reaction of the intermediate with the parent, probably to form benzodioxan "dimers." (The term dimer is used for convenience; the products are strictly bisdehydrodimers of dehydro-NADA (see "Discussion" and Fig. 11)) Rate constants of 5.9 x 10(5), 4.5 x 10(5), 2.8 x 10(4) and 3.5 x 10(4) M-1 s-1 were also obtained for decay of the second intermediate in the presence of cysteine, cysteamine, o-phenylenediamine, and p-aminophenol, respectively. By comparison with the UV-visible spectroscopic properties of the two-electron oxidized species derived from dehydro-NADA and from 1,2-dehydro-N-acetyldopa methyl ester, it is concluded that the transient intermediate exhibiting absorbance at 485 nm is the quinone methide tautomer of the o-quinone of dehydro-NADA. Sclerotization of insect cuticle is discussed in the light of these findings.  相似文献   

4.
We have studied the oxidation of microperoxidase-8 (MP-8) by H2O2 and the subsequent reaction of the intermediates with substrate by stopped-flow experiments. Oxidation of MP-8 by H2O2 gives two intermediates, I and II. The observed rate constant for the formation of I is linearly dependent on [H2O2] and exhibits a bell-shaped dependence on pH with pKa values of 8.90 and 10.60, which are attributed to the deprotonation of MP-bound H2O2 and H2O, respectively. The observed rate constant for the conversion of I to II is independent of [H2O2], but increases sharply at pH>9.0. The predominant forms of the intermediate at pH 7.0 and 10.7 are I and II, respectively. Addition of substrate to the intermediates at pH 9.0 gives rise to three distinct stages, corresponding to the three steps (in decreasing order of rate): I-->II*, II-->MP, and II*-->MP. The rates of these steps are all linearly dependent on the substrate concentration and each individual rate constant has been determined. Substrate reactivity at pH 10.7 covers over two orders of magnitude, ranging from 1.36 x 10(7) M(-1) s(-1) for 1-naphthol to 4.03 x 10(4) M(-1) s(-1) for ferrocyanide. The substrate reactivity is linearly correlated with its reduction potential, indicating that an electron transfer process is involved in the rate-limiting step.  相似文献   

5.
Hemopexin (HPX), serving as scavenger and transporter of toxic plasma heme, has been postulated to play a key role in the homeostasis of NO. Here, kinetics of HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO are reported. NO reacts reversibly with HPX-heme(II) yielding HPX-heme(II)-NO, according to the minimum reaction scheme: HPX-heme(II)+NO kon<-->koff HPX-heme(II)-NO values of kon, koff, and K (=kon/koff) are (6.3+/-0.3)x10(3)M-1s-1, (9.1+/-0.4)x10(-4)s-1, and (6.9+/-0.6)x10(6)M-1, respectively, at pH 7.0 and 10.0 degrees C. O2 reacts with HPX-heme(II)-NO yielding HPX-heme(III) and NO3-, by means of the ferric heme-bound peroxynitrite intermediate (HPX-heme(III)-N(O)OO), according to the minimum reaction scheme: HPX-heme(II)-NO+O2 hon<--> HPX-heme(III)-N(O)OO l-->HPX-heme(III)+NO3- the backward reaction rate is negligible. Values of hon and l are (2.4+/-0.3)x10(1)M-1s-1 and (1.4+/-0.2)x10(-3)s-1, respectively, at pH 7.0 and 10.0 degrees C. The decay of HPX-heme(III)-N(O)OO (i.e., l) is rate limiting. The HPX-heme(III)-N(O)OO intermediate has been characterized by optical absorption spectroscopy in the Soret region (lambdamax=409 nm and epsilon409=1.51x10(5)M-1cm-1). These results, representing the first kinetic evidence for HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO, might be predictive of transient (pseudo-enzymatic) function(s) of heme carriers.  相似文献   

6.
The reaction kinetics of the peroxidase activity of prostaglandin H synthase have been examined with 15-hydroperoxyeicosatetraenoic acid and hydrogen peroxide as substrates and tetramethylphenylenediamine as cosubstrate. The apparent Km and Vmax values for both hydroperoxides were found to increase linearly with the cosubstrate concentration. The overall reaction kinetics could be interpreted in terms of an initial reaction of the synthase with hydroperoxide to form an intermediate equivalent to horseradish peroxidase Compound I, followed by reduction of this intermediate by cosubstrate to regenerate the resting enzyme. The rate constants estimated for the generation of synthase Compound I were 7.1 X 10(7) M-1 s-1 with the lipid hydroperoxide and 9.1 X 10(4) M-1 s-1 with hydrogen peroxide. The rate constants estimated for the rate-determining step in the regeneration of resting enzyme by cosubstrate were 9.2 X 10(6) M-1 s-1 in the case of the reaction with lipid hydroperoxide and 3.5 X 10(6) M-1 s-1 in the case of reaction with hydrogen peroxide. The intrinsic affinities of the synthase peroxidase for substrate (Ks) were estimated to be on the order of 10(-8) M for lipid hydroperoxide and 10(-5) M for hydrogen peroxide. These affinities are quite similar to the reported affinities of the synthase for these hydroperoxides as activators of the cyclooxygenase. The peroxidase activity was found to be progressively inactivated during the peroxidase reaction. The rate of inactivation of the peroxidase was increased by increases in hydroperoxide level, and decreased by increases in peroxidase cosubstrate. The inactivation of the peroxidase appeared to occur by a hydroperoxide-dependent process, originating from synthase Compound I or Compound II.  相似文献   

7.
Stopped-flow techniques were used to investigate the kinetics of the formation of manganese peroxidase compound I (MnPI) and of the reactions of MnPI and manganese peroxidase compound II (MnPII) with p-cresol and MnII. All of the rate data were obtained from single turnover experiments under pseudo-first order conditions. In the presence of H2O2 the formation of MnPI is independent of pH over the range 3.12-8.29 with a second-order rate constant of (2.0 +/- 0.1) x 10(6) M-1 s-1. The activation energy for MnPI formation is 20 kJ mol-1. MnPI formation also occurs with organic peroxides such as peracetic acid, m-chloroperoxybenzoic acid, and p-nitroperoxybenzoic acid with second-order rate constants of 9.7 x 10(5), 9.5 x 10(4), and 5.9 x 10(4) M-1 s-1, respectively. The reactions of MnPI and MnPII with p-cresol strictly obeyed second-order kinetics. The second-order rate constant for the reaction of MnPII with p-cresol is extremely low, (9.5 +/- 0.5) M-1 s-1. Kinetic analysis of the reaction of MnII with MnPI and MnPII showed a binding interaction with the oxidized enzymes which led to saturation kinetics. The first-order dissociation rate constants for the reaction of MnII with MnPI and MnPII are (0.7 +/- 0.1) and (0.14 +/- 0.01) s-1, respectively, when the reaction is conducted in lactate buffer. Rate constants are considerably lower when the reactions are conducted in succinate buffer. Single turnover experiments confirmed that MnII serves as an obligatory substrate for MnPII and that both oxidized forms of the enzyme form productive complexes with MnII. Finally, these results suggest the alpha-hydroxy acids such as lactate facilitate the dissociation of MnIII from the enzyme.  相似文献   

8.
The mechanism of action of the flavoprotein D-aspartate oxidase (EC 1.4.3.1) has been investigated by steady-state and stopped flow kinetic studies using D-aspartate and O2 as substrates in 50 mM KPi, 0.3 mM EDTA, pH 7.4, 4 degrees C. Steady-state results indicate that a ternary complex containing enzyme, O2, and substrate (or product) is an obligatory intermediate in catalysis. The kinetic parameters are turnover number = 11.1 s-1, Km(D-Asp) = 2.2 x 10(-3) M, Km(O2) = 1.7 x 10(-4) M. Rapid reaction studies show that 1) the reductive half reaction is essentially irreversible with a maximum rate of reduction of 180 s-1; 2) the free reduced enzyme cannot be the species which is reoxidized during turnover since its reoxidation by oxygen (second order rate constant equal to 5.3 x 10(2) M-1 s-1) is too slow to be of relevance in catalysis; 3) reduced enzyme can bind a ligand rapidly and be reoxidized as a complex at a rate faster than that observed for the free reduced enzyme; 4) the rate of reoxidation of reduced enzyme by oxygen during turnover is dependent on both O2 and D-aspartate concentrations (second order rate constant of reaction between O2 and reduced enzyme-substrate complex equal to 6.2 x 10(4) M-1 s-1); and 5) the rate-limiting step in catalysis occurs after reoxidation of the enzyme and before its reduction in the following turnover. A mechanism involving reduction of enzyme by substrate, dissociation of product from reduced enzyme, binding of a second molecule of substrate to the reduced enzyme, and reoxidation of the reduced enzyme-substrate complex is proposed for the enzyme-catalyzed oxidation of D-aspartate.  相似文献   

9.
The Ca2+-binding kinetics of fura-2 and azo-1 were studied using temperature-jump relaxation methods. In 140 mM KCl at 20 degrees C, the association and dissociation rate constants for fura-2 were 6.02 x 10(8) M-1s-1 and 96.7 s-1, respectively. The fura-2 kinetics were insensitive to pH over the range 7.4 to 8.4. Azo-1 was studied in 140 mM KCl, at pH 7.4, at 10 degrees and 20 degrees C. At 10 degrees C, azo-1 exhibited association and dissociation rate constants of 1.43 x 10(8) M-1s-1 and 777.9 s-1, respectively; while at 20 degrees C, the corresponding values were 3.99 x 10(8) M-1s-1 and 1,177 s-1. The kinetic results demonstrate that fura-2 and azo-1 are well suited to monitoring rapid changes in intracellular [Ca2+].  相似文献   

10.
A green manganese desferrioxamine complex is rapidly formed at room temperature upon stirring freshly precipitated manganese dioxide in a solution of the ligand. Spectral studies and low-temperature ESR indicate that this compound, which has been previously described as a manganese(III) complex, is better characterized as containing tetravalent manganese. The complex appears to form oligomers in solution. The extinction coefficient at 635 nm is 137 +/- 6 M-1 cm-1 (per manganese) at pH 7.8 and 88 +/- 4 M-1 s-1 at pH 6.6 after purification by chromatography. The superoxide dismutase activity was measured and compared to that of mononuclear manganese(III) 1,4,8,11-tetraazacyclodecane (cyclam). The catalytic rate constants for superoxide dismutase activity are 1.7 x 10(6) M-1 s-1 and 2.9 x 10(6) M-1 s-1 for the desferrioxamine and the cyclam complexes, respectively.  相似文献   

11.
The reaction of human myeloperoxidase with its product, hypochlorous acid was investigated using both rapid-scan spectrophotometry and the stopped-flow technique. In the reaction of myeloperoxidase with hypochlorous acid a primary compound is found with properties similar to that of compound I and which is converted into compound II. The primary reaction is strongly pH-dependent. At pH 7.2 the reaction is too fast to be measured but at higher pH values it is possible to determine the apparent second-order rate constant. Its value decreases to about 2 x 10(7) M-1.s-1 at pH 8.3 and to 2.3 (+/- 0.4) x 10(6) M-1.s-1 at pH 9.2, respectively. The dissociation constant for the formation of the primary compound is 25.7 (+/- 15.3) microM at pH 9.2 and about 2.5 microM at pH 8.3. The apparent second-order rate constant for the formation of compound II is hardly affected by pH and varies between 2 to 5 x 10(4) M-1.s-1 at pH 10.2 and pH 8.3, respectively. Reaction of myeloperoxidase with hypochlorous acid also resulted in irreversible partial bleaching of the chromophore. Chloride, which is a substrate of the enzyme not only protects myeloperoxidase against bleaching by hypochlorous acid but also competitively inhibits the binding of hypochlorous acid to myeloperoxidase, a process which also has been observed in the reaction with hydrogen peroxide. It is concluded that hypochlorous acid binds at the heme iron to form compound I.  相似文献   

12.
The acid-basic properties of ellipticine have been re-estimated. The apparent pK of protonation at 3 microM drug concentration is 7.4 +/- 0.1. The ellipticine free base (at pH 9, I = 25 mM) intercalates into calf-thymus DNA with an affinity constant of 3.3 +/- 0.2 X 10(5) M-1, and a number of binding sites per phosphate of 0.23. The ellipticinium cation (pH 5, I = 25 mM) binds also to DNA with a constant of 8.3 +/- 0.2 x 10(5) M-1 and at a number of binding sites (n = 0.19). It is postulated that the binding of the drug to DNA at pH 9 is driven by hydrophobic and/or dipolar effects. Even at pH 5, where ellipticine exists as a cation, it is thought that the hydrophobic interaction is the main contribution to binding. The neutral and cationic forms share common binding within DNA sites but yield to structurally different complexes. The free base has 0.04 additional specific binding sites per phosphate. As determined from temperature-jump experiments, the second-order rate constant of the binding of the free base (pH 9) is 3.4 x 10(7) M-1 s-1 and the residence time of the base within the DNA is 8 ms. The rate constant for the binding of the ellipticinium cation is 9.8 x 10(7) M-1 s-1 when it is assumed that drug attachment occurs via a pathway in which the formation of an intermediate ionic complex is not involved (competitive pathway).  相似文献   

13.
Kinetics of CO combination with and dissociation from isomer III, (alpha 1CO beta 1CO)alpha 2 beta 2 or alpha 1 beta 1 (alpha 2CO beta 2CO), and Hb Rothschild have been studied using the double mixing and microperoxidase methods. Isomer III was prepared in a manner so that it was the only reactive species in the reaction mixture. The biphasic reaction time course in both the "on" and "off" reactions of isomer III and the CO combination reaction of Hb Rothschild are attributed to slow relaxation between the fast and slow CO-reacting species in the two proteins: isomer III: l'f = 6 x 10(6) M-1 s-1, l'dimer = 1.7 x 10(6) M-1 s-1, l's = 2.2 x 10(5) M-1 s-1, lf = 0.15 s-1, ls = 0.01 s-1; Hb Rothschild: l'f = 2.8 x 10(6) M-1 s-1; l's = 2.7 x 10(5) M-1 s-1.  相似文献   

14.
The interplay of different proteases and glycosaminoglycans is able to modulate the activity of the enzymes and to affect their structures. Human plasma kallikrein (huPK) is a proteolytic enzyme involved in intrinsic blood clotting, the kallikrein-kinin system and fibrinolysis. We investigated the effect of heparin on the action, inhibition and secondary structure of huPK. The catalytic efficiency for the hydrolysis of substrates by huPK was determined by Michaelis-Menten kinetic plots: 5.12x10(4) M-1 s-1 for acetyl-Phe-Arg-p-nitroanilide, 1.40x10(5) M-1 s-1 for H-D-Pro-Phe-Arg-p-nitroanilide, 2.25x10(4) M-1 s-1 for Abz-Gly-Phe-Ser-Pro-Phe-Arg-Ser-Ser-Arg-Gln-EDDnp, 4.24x10(2)M-1 s-1 for factor XII and 5.58x10(2) M-1 s-1 for plasminogen. Heparin reduced the hydrolysis of synthetic substrates (by 2.0-fold), but enhanced factor XII and plasminogen hydrolysis (7.7- and 1.4-fold, respectively). The second-order rate constants for inhibition of huPK by antithrombin and C1-inhibitor were 2.40x10(2) M-1 s-1 and 1.70x10(4) M-1 s-1, respectively. Heparin improved the inhibition of huPK by these inhibitors (3.4- and 1.4-fold). Despite the fact that huPK was able to bind to a heparin-Sepharose matrix, its secondary structure was not modified by heparin, as monitored by circular dichroism. These actions may have a function in the control or maintenance of some pathophysiological processes in which huPK participates.  相似文献   

15.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

16.
Protein methylesterase (PME) amino acid composition and substrate specificity towards methylated normal and deamidated protein substrates were investigated. The enzyme contained 23% acidic and 5% basic residues. These values are consistent with a pI of 4.45. The product formed from methylated protein by PME was confirmed as methanol by h.p.l.c. The kcat. and Km values for several methylated protein substrates ranged from 20 x 10(-6) to 560 x 10(-6) s-1 and from 0.5 to 64 microM respectively. However, the kcat./Km ratios ranged within one order of magnitude from 11 to 52 M-1.s-1. Results with the irreversible cysteine-proteinase inhibitor E-64 suggested that these low values were in part due to the fact that only one out of 25 molecules in the PME preparations was enzymically active. When PME was incubated with methylated normal and deamidated calmodulin, the enzyme hydrolysed the latter substrate at a higher rate. The Km and kcat. for methylated normal calmodulin were 0.9 microM and 31 x 10(-6) s-1, whereas for methylated deamidated calmodulin values of 1.6 microM and 188 x 10(-6) s-1 were obtained. The kcat./Km ratios for methylated normal and deamidated calmodulin were 34 and 118 M-1.s-1 respectively. By contrast, results with methylated adrenocorticotropic hormone (ACTH) substrates indicated that the main difference between native and deamidated substrates resides in the Km rather than the kcat. The Km for methylated deamidated ACTH was 5-fold lower than that for methylated native ACTH. The kcat./Km ratios for methylated normal and deamidated ACTH were 43 and 185 M-1.s-1 respectively. These results indicate that PME recognizes native and deamidated methylated substrates as two different entities. This suggests that the methyl groups on native calmodulin and ACTH substrates may not be on the same amino acid residues as those on deamidated calmodulin and ACTH substrates.  相似文献   

17.
Hydroxyl radicals abstract hydrogen atoms from glycerol-2-phosphate with a specific rate constant of (7.0 +/- 1.5) x 10(8) M-1s-1 forming the beta-phospho radical as the major product. At physiological pH this radical undergoes a beta-phosphate elimination with a rate constant less than or equal to 1 x 10(3) s-1. The beta-phospho radical reacts with Cu(I)-phenanthroline to produce an unstable transient with a metal-carbon sigma-bond which has an absorbance similar to that of the cuprous phenanthroline complex in the visible region. This intermediate decomposes via a beta-elimination of phosphate with a rate constant of (1.0 +/- 1.5) x 10(4) s-1, which was independent of the acidity in the pH range 4-9.  相似文献   

18.
The putative intermediate dienol (2) in the steroid isomerase (KSI) catalyzed conversion of 5-androstene-3,17-dione (1) to 4-androstene-3,17-dione (3) has been independently generated and tested as a substrate for KSI. At pH 7, dienol 2 is converted by KSI to a mixture of 1 (46%) and 3 (54%). The apparent second-order rate constant for reaction of 2 with KSI to produce 3 (kappa cat/Km = 2.3 x 10(8) M-1 s-1) is similar to that for reaction of 1 with KSI (kappa cat/Km = 2.1 x 10(8) M-1 s-1), demonstrating that 2 is kinetically competent. Isomerization of 1 by KSI in D2O gives only 5% of solvent deuterium incorporated into the product 3. When 2 reacts with KSI in D2O, and the product 3 is isolated (from direct reaction of 2 and from subsequent conversion of the 1 initially formed), ca. 80 atom % deuterium is located at C-6 beta, confirming that protonation of the dienol by KSI occurs at the same face as the proton transfer in the KSI catalyzed reaction of 1 to 3.  相似文献   

19.
Three hemoglobins have been isolated from the symbiont-harboring gill of the bivalve mollusc Lucina pectinata. Oxyhemoglobin I (Hb I), which may be called sulfide-reactive hemoglobin, reacts with hydrogen sulfide to form ferric hemoglobin sulfide in a reaction that may proceed by nucleophilic displacement of bound superoxide anion by hydrosulfide anion. Hemoglobins II and II, called oxygen-reactive hemoglobins, remain oxygenated in the presence of hydrogen sulfide. Hemoglobin I is monomeric; Hb II and Hb III self-associate in a concentration-dependent manner and form a tetramer when mixed. Oxygen binding is not cooperative. Oxygen affinities are all nearly the same, P50 = 0.1 to 0.2 Torr, and are independent of pH. Combination of Hb I with oxygen is fast; k'on = (estimated) 100-200 x 10(6) M-1 s-1. Combination of Hb II and Hb III with oxygen is slow: k'on = 0.4 and 0.3 x 10(6) M-1 s-1, respectively. Dissociation of oxygen from Hb I is fast relative to myoglobin: koff = 61 s-1. Dissociation from Hb II and Hb III is slow: koff = 0.11 and 0.08 s-1, respectively. These large differences in rates of reaction together with differences in the reactions of carbon monoxide suggest differences in configuration of the distal heme pocket. The fast reactions of Hb I are comparable to those of hemoglobins that lack distal histidine residues. Slow dissociation of oxygen from Hb II and Hb III suggest that a distal residue may interact strongly with the bound ligand. We infer that Hb I may facilitate delivery of hydrogen sulfide to the chemoautotrophic bacterial symbiont and Hb II and Hb III may facilitate delivery of oxygen. The midpoint oxidation-reduction potential of the ferrous/ferric couple of Hb I, 103 +/- 8 mV, was independent of pH. Potentials of Hb II and Hb III were pH-dependent. At neutral pH all three hemoglobins have similar midpoint potentials. The rate constant for combination of ferric Hb I with hydrogen sulfide increases 3000-fold from pH 10.5 to 5.5, with apparent pK 7.0, suggesting that undissociated hydrogen sulfide is the attacking ligand. At the acid limit combination of ferric Hb I with hydrogen sulfide, k'on = 2.3 x 10(5) M-1 s-1, is 40-fold faster than combination with ferric Hb II or myoglobin.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
Structural and functional properties of alpha-protease nexin I (alpha-PNI) expressed in Chinese hamster ovary cells were studied. All three cysteines were in the reduced form, showing that the potential disulfide bridge between residues Cys117 and Cys131 was not formed. Heparin association rate enhancements were from ka = 8.3 x 10(5) to 0.7-1.6 x 10(9) M-1 s-1 for the interaction of PNI with thrombin, from ka = 5.1 x 10(3) to 3.5 x 10(5) M-1 s-1 for interaction with Factor Xa, and from ka = 2.2 x 10(6) to 1.0 x 10(7) M-1 s-1 for interaction with trypsin; there was no rate enhancement of the plasmin interaction (ka = 1.0 x 10(5) M-1 s-1). The minimal heparin pentasaccharide had no effect on these interactions. Cleavage of the reactive center loop of PNI by three different proteases gave the typical stressed to relaxed change in thermal stability, but unlike with antithrombin III, there was no loss of heparin affinity. A similar difference from antithrombin was that PNI-thrombin complexes retained normal heparin affinity. These results are compatible with a role for protease nexin I as a cell-associated thrombin inhibitor that remains bound to the cell surface even after complexing with the protease, as compared with the role of antithrombin III as a circulating inhibitor of thrombin that becomes activated on binding to the microvasculature and is released on complex formation.  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