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.
Stopped-flow techniques were utilized to investigate the kinetics of the reaction of lignin peroxidase compounds I and II (LiPI and LiPII) with veratryl alcohol (VA). All rate data were collected from single turnover experiments under pseudo first-order conditions. The reaction of LiPI with VA strictly obeys second-order kinetics over the pH range 2.72-5.25 as demonstrated by linear plots of the pseudo first-order rate constants versus concentrations of VA. The second-order rate constants are strongly dependent on pH and range from 2.62 x 10(6) M-1 s-1 (pH 2.72) to 1.45 x 10(4) M-1 s-1 (pH 5.25). The reaction of LiPII and VA exhibits saturation behavior when the observed pseudo first-order rate constants are plotted against VA concentrations. The saturation phenomenon is quantitatively explained by the formation of a 1:1 LiPII-substrate complex. Results of kinetic and rapid scan spectral analyses exclude the formation of a LiPII-VA cation radical complex. The first-order dissociation rate constant and the equilibrium dissociation constant for the LiPII reaction are also pH dependent. Binding of VA to LiPII is controlled by a heme-linked ionizable group of pKa approximately 4.2. The pH profiles of the second-order rate constants for the LiPI reaction and of the first-order dissociation constants for the LiPII reaction both demonstrate two pKa values at approximately 3.0 and approximately 4.2. Protonated oxidized enzyme intermediates are most active, suggesting that only electron transfer, not proton uptake from the reducing substrate, occurs at the enzyme active site. These results are consistent with the one-electron oxidation of VA to an aryl cation radical by LiPI and LiPII.  相似文献   

2.
The effect of temperature on the adsorption of metanil yellow (MY) (acidic) and methylene blue (MB) (basic) by poplar sawdust was investigated. In addition, the amounts of NaHCO(3), Na(2)CO(3), NaOH and C(2)H(5)ONa adsorbed by 1g of poplar sawdust to determine its surface acidity were also determined. Kinetical data obtained at different temperatures (293 K, 313 K and 333 K) for the adsorption of each dyestuff by poplar sawdust were applied to the pseudo first-order, the pseudo second-order and the intraparticle diffusion equations, and the rate constants of first-order adsorption (k(1)), the rate constants of second-order adsorption (k(2)) and intraparticle diffusion rate constants (k(p)) at these temperatures were calculated, respectively. In addition, isothermal data obtained at different temperatures (293 K, 313 K and 333 K) for the adsorption of each dyestuff by poplar sawdust were applied to thermodynamical equations, and thermodynamical parameters (DeltaG, DeltaH and DeltaS) were also calculated.  相似文献   

3.
Microbial transformation rate constants were determined for seven amides in natural pond water. A second-order mathematical rate expression served as the model for describing the microbial transformation. Also investigated was the relationship between the infrared spectra and the second-order rate constants for these amides. Second-order rate constants (k2) ranged from a low of 2.0 X 10(-14) to a high of 1.1 X 10(-9) liters organism-1 h-1 for niclosamide (2',5-dichloro-4'-nitrosalicylanilide) and propachlor (2-chloro-N-isopropylacetanilide), respectively. The mechanism of degradation (i.e., microbially mediated hydrolysis) of the amides was consistent with that of other organic chemicals previously studied in a variety of natural waters. Preliminary investigations indicate that temporal variations in measured second-order rate constants are small. A simple linear regression of the infrared carbonyl-stretching frequency with log K2 gave a correlation coefficient (r2) of 0.962.  相似文献   

4.
Microbial transformation rate constants were determined for seven amides in natural pond water. A second-order mathematical rate expression served as the model for describing the microbial transformation. Also investigated was the relationship between the infrared spectra and the second-order rate constants for these amides. Second-order rate constants (k2) ranged from a low of 2.0 X 10(-14) to a high of 1.1 X 10(-9) liters organism-1 h-1 for niclosamide (2',5-dichloro-4'-nitrosalicylanilide) and propachlor (2-chloro-N-isopropylacetanilide), respectively. The mechanism of degradation (i.e., microbially mediated hydrolysis) of the amides was consistent with that of other organic chemicals previously studied in a variety of natural waters. Preliminary investigations indicate that temporal variations in measured second-order rate constants are small. A simple linear regression of the infrared carbonyl-stretching frequency with log K2 gave a correlation coefficient (r2) of 0.962.  相似文献   

5.
J Steyaert  L Wyns  P Stanssens 《Biochemistry》1991,30(35):8661-8665
We report on the effect of the viscogenic agents glycerol and ficoll on the RNase T1 catalyzed turnover of GpA, GpC, GpU, and Torula yeast RNA. For wild-type enzyme, we find that the kcat/Km values for the transesterification of GpC and GpA as well as for the cleavage of RNA are inversely proportional to the relative viscosity of glycerol-containing buffers; no such effect is observed for the conversion of GpU to cGMP and U. The second-order rate constants for His40Ala and Glu46Ala RNase T1, two mutants with a drastically reduced kcat/km ratio, are independent of the microviscosity, indicating that glycerol does not affect the intrinsic kinetic parameters. Consistent with the notion that molecular diffusion rates are unaffected by polymeric viscogens, addition of ficoll has no effect on the kcat/Km for GpC transesterification by wild-type enzyme. The data indicate that the second-order rate constants for GpC, GpA, and Torula yeast RNA are at least partly limited by the diffusion-controlled association rate of substrate and active site; RNase T1 obeys Briggs-Haldane kinetics for these substrates (Km greater than Ks). Calculations suggest that the equilibrium dissociation constants (Ks) for the various GpN-wild-type enzyme complexes are virtually independent of N whereas the measured kcat values follow the order GpC greater than GpA greater than GpU. This is also revealed by the steady-state kinetic parameters of Tyr38Phe and His40Ala RNase T1, two mutants that follow simple Michaelis-Menten kinetics because of a dramatically reduced kcat value (i.e., Km = Ks).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Summary The second-order rate constants characterizing the association of potential-sensing dyes of the cyanine, merocyanine, and oxonol classes with glycerylmonooleate suspensions, azolectin vesicles, or submitochondrial particles have been measured and the implications for redistribution type mechanisms proposed to explain the potential-dependent optical signals of these probes considered. The second-order rate constants obtained for the cyanines and oxonols are compatible with microsecond probe response times only on the assumption that a high local dye concentration exists in the aqueous phase immediately adjacent to the membrane surface. Calculations based on a surface charge density induced by a bias potential suggest that the necessary local concentration cannot be attained by a diffusion polarization mechanism. A model based on the rapid recombination of ejected dye with the membrane bilayer seems capable of explaining microsecond probe response times in systems where the potential is rapidly changing polarity; calculations suggest that an ejected dye molecule would not diffuse out of an unstirred layer of 100 microns thickness on a millisecond time scale. Microsecond probe responses are also compatible with a first-order potential-dependent dye ejection from the membrane with no rapid recombination when the potential is not changing polarity. The apparent first-order rate constants describing the interaction of merocanine M-540 with a glycerylmonooleate suspension are independent of dye concentration; the reaction may be diffusion limited. The high local dye concentration need not be met in this case for a mechanism based on the transfer of dye onto the membrane from the aqueous phase to describe the microsecond signals of this dye, but other mechanisms have been proposed to explain such signals. The mechanism leading to potentialdependent signals from optical probes appear to differ substantially between suspensions of energy-transducing biological membranes and those involving excitable membranes such as the squid giant axon or model black lipid membranes.  相似文献   

7.
The reliability of second-order rate constants for assessing microbial degradation kinetics in natural waters was examined by using three compounds that undergo hydrolytic degradation. The butoxyethyl ester of 2,4-dichlorophenoxyacetic acid was studied in water samples from 31 sites, malathion was examined in water from 14 sites, and chlorpropham was studied in samples from 11 sites. The coefficient of variation for rate constants for each compound was less than 65% over all sites. Additional studies indicated that the rate conformed to second-order kinetics; that is, the rate was proportional to both bacterial and xenobiotic concentrations.  相似文献   

8.
Glial-derived neurite-promoting factor was found to be a slow-binding inhibitor of trypsin, urokinase, and thrombin. The kinetic mechanism of the inhibition differs among the three proteases. With trypsin and urokinase, an initial protease-factor complex formed which isomerized to a tighter complex. For thrombin, however, no initial complex was kinetically observed. The dissociation constants of the equilibrium complexes of the factor with trypsin, urokinase, and thrombin were 17, 280, and 18 pM, respectively, and the apparent second-order rate constants for the interaction of the factor with these enzymes were, respectively, 4.7 X 10(6), 1.2 X 10(5), and 2.1 X 10(6) M-1S-1. Heparin increased the rate at which the factor reacted with thrombin by over 40-fold to 8.9 X 10(7) M-1S-1 and decreased the dissociation constant of the complex by over 80-fold to 0.3 pM. The values obtained for the apparent second-order rate constants when compared with the kinetics of neurite induction by the factor indicate that the neurite-promoting activity of the factor is not due to the inhibition of urokinase but could be due to the inhibition of an enzyme with a specificity similar to that of thrombin or trypsin. Comparison of the values of the apparent second-order rate constants obtained for the factor with those obtained for protease nexin suggests that these two molecules are very similar in their inhibitory properties.  相似文献   

9.
We investigated the effect of preincubation of environmental waters amended with inorganic nutrients (nitrogen, phosphorus, and traces of iron and magnesium) on the kinetics of the microbial transformation of phenol, propanil, propyl ester of (2,4-dichlorophenoxy)acetic acid, methyl parathion, Ronnel, and methoxychlor in pond and river waters. No effect on the second-order rate constants for these compounds was observed, although there was an increase in the bacterial populations and the pseudo-first-order rate constants. The use of nutrient-amended waters could be a useful tool for estimating second-order rate constants for an expanded number of compounds. This technique would provide a larger data base for predicting the behavior of xenobiotic compounds in the environment by using currently available mathematical models.  相似文献   

10.
DNA reassociation kinetics using the phenol emulsion reassociation technique (PERT) [Kohne, D. E., Levison, S. A. & Byers, M. J. (1977) Biochemistry 16 , 5329–5341] has been investigated at high DNA concentrations using an endonuclease S1 assay of reaction progress. Apparent second-order rate constants fall on two intersecting straight lines when presented as a function of DNA concentrations on a log–log plot. In the low DNA concentration range, the rate constants drop about 10-fold when concentration increases 1000-fold. In the high DNA concentration range, the rate constants drop more than 10-fold when concentration increases 10-fold. The slopes of these lines are the same in different solvents and at different temperatures. The intersection between the lines occurs when the available catalytic surface is saturated. At high DNA concentrations, high-complexity heterologous denatured DNA apparently competes 2–4 times better for the surface than homologous DNA because it does not participate in a reassociation reaction. Native and partially native DNA molecules cannot compete with single-stranded DNA for a saturated surface. At high DNA concentrations, reactions using PERT become dependent on the single-strand DNA length. Increasing length lowers reassociation rates.  相似文献   

11.
The chemical properties of the nitro radical of four types of nitroheterocyclic compounds, nitrofurans, 2-nitroimidazoles, 4(5)-nitroimidazoles, 5-nitroimidazoles, having radiosensitizing and cytotoxic properties, have been studied by pulse radiolysis. The acid-base equilibria involving the nitro radical, the imidazole ring and some residues on the heterocycle have been determined. The pH-dependence of the rate of the disproportionation reaction of the nitro radical have been extensively studied. While the nitro radical derived from nitrofurans, 4- and 5-nitroimidazoles had a second-order decay, those of the 2-nitroimidazoles were found to decay through simultaneous first-order and second-order processes. Intrinsic second-order rate constants of the decay of the radical species in its various acidic and basic forms, could be determined. The intrinsic rate constants that determine the overall decay rates in the physiologically important 6 to 7.5 pH-range could be related to the one-electron redox potential E7(1). The implication of such chemical properties to enzyme-catalyzed reduction processes and to the mechanisms of radiosensitization and cytotoxicity of nitroheterocyclic compounds are briefly discussed. Pharmacological properties such as in vitro radiosensitization efficiency or metabolic reduction rates could be related to two of the nitro radical intrinsic disproportionation rates.  相似文献   

12.
D Herschlag  W P Jencks 《Biochemistry》1990,29(21):5172-5179
The second-order rate constants for reaction of the Mg2+ complexes of phosphorylated pyridine monoanions with Mg(OH)+ are 10(4)-10(6)-fold larger than the second-order rate constants for their reaction with water (25 degrees C, ionic strength 1.5). Of the 10(6)-fold rate enhancement with the phosphorylated 4-morpholinopyridine/Mg2 complex, approximately 10(4)-fold is attributed to the greater nucleophilicity of Mg(OH)+ compared with water. The remaining catalysis of approximately 10(2)-fold is attributed to induced intramolecularity from positioning of the hydroxide ion and phosphoryl group by the Mg2+ ions. This reaction may provide a model for the role of a metal ion in increasing the concentration of the anions of enolpyruvate and serine and holding the nucleophile in the correct position for phosphoryl transfer in the reactions catalyzed by pyruvate kinase and alkaline phosphatase, for example. Some mechanisms that can provide catalysis of phosphoryl transfer through a metaphosphate-like transition state are reviewed briefly.  相似文献   

13.
The reaction between cytochrome c1 and cytochrome c   总被引:3,自引:0,他引:3  
The kinetics of electron transfer between the isolated enzymes of cytochrome c1 and cytochrome c have been investigated using the stopped-flow technique. The reaction between ferrocytochrome c1 and ferricytochrome c is fast; the second-order rate constant (k1) is 3.0 . 10(7) M-1 . s-1 at low ionic strength (I = 223 mM, 10 degrees C). The value of this rate constant decreases to 1.8 . 10(5) M-1 . s-1 upon increasing the ionic strength to 1.13 M. The ionic strength dependence of the electron transfer between cytochrome c1 and cytochrome c implies the involvement of electrostatic interactions in the reaction between both cytochromes. In addition to a general influence of ionic strength, specific anion effects are found for phosphate, chloride and morpholinosulphonate. These anions appear to inhibit the reaction between cytochrome c1 and cytochrome c by binding of these anions to the cytochrome c molecule. Such a phenomenon is not observed for cacodylate. At an ionic strength of 1.02 M, the second-order rate constants for the reaction between ferrocytochrome c1 and ferricytochrome c and the reverse reaction are k1 = 2.4 . 10(5) M-1 . s-1 and k-1 = 3.3 . 10(5) M-1 . s-1, respectively (450 mM potassium phosphate, pH 7.0, 1% Tween 20, 10 degrees C). The 'equilibrium' constant calculated from the rate constants (0.73) is equal to the constant determined from equilibrium studies. Moreover, it is shown that at this ionic strength, the concentrations of intermediary complexes are very low and that the value of the equilibrium constant is independent of ionic strength. These data can be fitted into the following simple reaction scheme: cytochrome c2+1 + cytochrome c3+ in equilibrium or formed from cytochrome c3+1 + cytochrome c2+.  相似文献   

14.
The homogeneous recombinant mammalian protein tyrosine phosphatase 1B (PTP1B) and Yersinia protein tyrosine phosphatase (PTPase) are inactivated by a series of low-molecular-weight S-nitrosothiols. These compounds exhibited different inhibitory activities in a time- and concentration-dependent manner with second-order rate constants (k(inact)/K(I)) ranging from 37 to 113 M(-1) min(-1) against mammalian PTP1B and from 66 to 613 M(-1) min(-1) against Yersinia PTPase. Furthermore, the inactivation of Yersinia PTPase by S-nitrosylated protein:S-nitroso human serum albumin was investigated. Both single-S-nitrosylated and poly-S-nitrosylated human serum albumin show good inhibitory ability to Yersinia PTPase. The second-order rate constants are 472 and 1188 M(-1) min(-1), respectively. This result indicates a possibility that S-nitrosylated albumin in vivo may function as an inhibitor for a variety of cysteine-dependent enzymes.  相似文献   

15.
High-molecular-weight penicillin-binding proteins (PBPs) are essential integral membrane proteins of the bacterial cytoplasmic membrane responsible for biosynthesis of peptidoglycan. They are the targets of antibacterial β-lactam drugs, including penicillins, cephalosporins, and carbapenems. β-Lactams covalently acylate the active sites of the PBP transpeptidase domains. Because β-lactams are time-dependent inhibitors, quantitative assessment of the inhibitory activity of these compounds ideally involves measurement of their second-order acylation rate constants. We previously described a fluorescence anisotropy-based assay to measure these rate constants for soluble constructs of PBP3 (Anal. Biochem. 439 (2013) 37–43). Here we report the expression and purification of a soluble construct of Pseudomonas aeruginosa PBP2 as a fusion protein with NusA. This soluble PBP2 was used to measure second-order acylation rate constants with the fluorescence anisotropy assay. Measurements were obtained for mecillinam, which reacts specifically with PBP2, and for several carbapenems. The assay also revealed that PBP2 slowly hydrolyzed mecillinam and was used to measure the rate constant for this deacylation reaction.  相似文献   

16.
The kinetics of release of Al3+ from human serum dialuminum transferrin (Al2Tf) to citrate were investigated at 37 degrees C, pH 7.4, mu = 0.7 M, by difference UV spectrophotometry. The two metal-binding sites are not identical but behave in a kinetically similar manner to give apparent second-order rate constants of 0.60 and 0.38 M-1 s-1, respectively, for release of the first Al3+ from Al2Tf. The rate constants for release of the second metal ion from the monoaluminum transferrins are 0.27 and 0.12 M-1 s-1. The kinetic scheme for release of A13+ from Al2Tf is therefore similar to that for release of Fe3+ from Fe2Tf, but the rate of constants for metal ion release are between two and four orders of magnitude larger.  相似文献   

17.
Autoactivation of the proenzyme form of a subunit of the first component (C1r) was performed in the presence and absence of diisopropyl fluorophosphate (DFP). The time-course of autoactivation of zymogen C1r followed a sigmoidal curve and was accelerated by addition of the enzyme C1r and by increasing the concentration of C1r, suggesting that autoactivation of C1r consists of two intermolecular reactions, i.e. zymogen(C1r)- and enzyme(C1r)-catalyzed reactions. In the presence of 10 mM DFP, the enzyme-catalyzed autoactivation of C1r was completely inhibited, while the zymogen-catalyzed autoactivation still proceeded depending upon C1r concentration. These results suggested that the zymogen-catalyzed autoactivation of C1r is a DFP-insensitive second-order reaction and is mediated by an active site generated in a single chain C1r through a conformational change (Kassahara et al. (1982) FEBS lett. 141, 128-131). Based on these results, a possible reaction process of autoactivation of C1r was proposed, as follows: (formula; see text) where C1r represents a conformational isomer which catalyzes the autoactivation of C1r, and the rate constants, k2 and k3, are of second-order. Utilizing a computer, we simulated the autoactivation of C1r and found the above scheme to be a reasonable model of C1r autoactivation. Evidence which supports the formation of a conformational isomer of C1r, C1r, as an intermediate in its autoactivation was also obtained by a surface radiolabeling method.  相似文献   

18.
We have determined the rate constants of inactivation of factor Xa and thrombin by antithrombin III/heparin during the process of prothrombin activation. The second-order rate constant of inhibition of factor Xa alone by antithrombin III as determined by using the synthetic peptide substrate S-2337 was found to be 1.1 X 10(6) M-1 min-1. Factor Xa in prothrombin activation mixtures that contained prothrombin, and either saturating amounts of factor Va or phospholipid (20 mol % dioleoylphosphatidylserine/80 mol % dioleoylphosphatidylcholine, 10 microM), was inhibited by antithrombin III with a second-order rate constant that was essentially the same: 1.2 X 10(6) M-1 min-1. When both factor Va and phospholipid were present during prothrombin activation, factor Xa inhibition by antithrombin III was reduced about 10-fold, with a second-order rate constant of 1.3 X 10(5) M-1 min-1. Factor Xa in the prothrombin activation mixture that contained both factor Va and phospholipid was even more protected from inhibition by the antithrombin III-heparin complex. The first-order rate constants of these reactions at 200 nM antithrombin III and normalized to heparin at 1 microgram/mL were 0.33 and 9.5 min-1 in the presence and absence of factor Va and phospholipid, respectively. When the prothrombin concentration was varied widely around the Km for prothrombin, this had no effect on the first-order rate constants of inhibition. It is our conclusion that factor Xa when acting in prothrombinase on prothrombin is profoundly protected from inhibition by antithrombin III in the absence as well as in the presence of heparin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The reduction of single-site haem and copper redox proteins by ascorbic acid was studied as a function of pH. Evidence is presented that indicates that the double-deprotonated ascorbate anion, ascorbate2-, is the reducing agent, and the pH-independent second-order rate constants for reduction by this species are given. Investigation of the temperature dependences of these rate constants have yielded the values of the activation parameters (delta H++ and delta S++) for reduction. These values, together with ligand-replacement studies, suggest that ascorbate2- acts as an outer-sphere reductant for these proteins. Reasons to account for the apparent inability of ascorbic acid to reduce the alkaline conformer of mammalian ferricytochrome c are suggested.  相似文献   

20.
C Persson  E G Wagner    K Nordstrm 《The EMBO journal》1990,9(11):3777-3785
The replication frequency of plasmid R1 is determined by the availability of the initiator protein RepA. Synthesis of RepA is negatively controlled by an antisense RNA, CopA, which forms a duplex with the upstream region of the RepA mRNA, CopT. We have previously shown that the in vitro formation of the CopA-CopT duplex follows second-order kinetics and occurs in at least two steps. The first step is the formation of a transient (kissing) complex, which is subsequently converted to a persistent duplex. Here, we investigate the details of the reaction scheme and determine the rate constants of the pathway from the free RNAs to the complete duplex. Using a shortened CopA RNA (CopI) we have been able to determine the association and dissociation rate constants (k1,k-1) for the kissing complex (which are inferred to be the same for CopI-T and CopA-T), and measured the hybridization rate constant k2 (for CopA-T k2 is at least 1000-fold greater than for CopI-T). The analysis of CopA derivatives of mutant and wild-type origin shows that the rate of formation of the kissing complex is rate-limiting for the overall pairing reaction between CopA and CopT, both in vitro and in vivo. The biological implications of the kinetically irreversible RNA-RNA binding reaction scheme are discussed.  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