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.
Saccharomyces cerevisiae phosphoenolpyruvate (PEP) carboxykinase catalyzes the reversible formation of oxaloacetate and adenosine triphosphate from PEP, adenosine diphosphate and carbon dioxide, and uses Mn(2+) as the activating metal ion. Comparison with the crystalline structure of homologous Escherichia coli PEP carboxykinase [Tari et al. Nature Struct. Biol. 4 (1997) 990-994] shows that Lys(213) is one of the ligands to Mn(2+) at the enzyme active site. Coordination of Mn(2+) to a lysyl residue is infrequent and suggests a low pK(a) value for the epsilon-NH(2) group of Lys(213). In this work, we evaluate the role of neighboring Phe(416) in contributing to provide a low polarity microenvironment suitable to keep the epsilon-NH(2) of Lys(213) in the unprotonated form. Mutation Phe416Tyr shows that the introduction of a hydroxyl group in the lateral chain of the residue produces a substantial loss in the enzyme affinity for Mn(2+), suggesting an increase of the pK(a) of Lys(213). A study of the effect of pH on K(m) for Mn(2+) indicate that the affinity of recombinant wild type enzyme for the metal ion is dependent on deprotonation of a group with pK(a) of 7.1+/-0.2, compatible with the low pK(a) expected for Lys(213). This pK(a) value increases at least 1.5 pH units upon Phe416Tyr mutation, in agreement with the expected effect of an increase in the polarity of Lys(213) microenvironment. Theoretical calculations of the pK(a) of Lys(213) indicate a value of 6.5+/-0.9, and it increases to 8.2+/-1.6 upon Phe416Tyr mutation. Additionally, mutation Phe416Tyr causes a loss of 1.3 kcal mol(-1) in the affinity of the enzyme for PEP, an effect perhaps related to the close proximity of Phe(416) to Arg(70), a residue previously shown to be important for PEP binding.  相似文献   

2.
Inorganic polyphosphate (poly(P)) is a linear polymer that has been found in every organism so far examined. To elucidate the functions of poly(P) in the regulation of gene expression, the level of cellular poly(P) in Escherichia coli was reduced to a barely detectable concentration by overproduction of exopolyphosphatase (exopoly(P)ase) with a plasmid encoding yeast exopoly(P)ase (Shiba et al., Proc. Natl. Acad. Sci. USA 94 (1997) 11210-11215). It was found that exopoly(P)ase-overproducing cells were more sensitive to UV or mitomycin C (MMC) than were control cells. Poly(P) accumulation was observed after treatment with MMC, whereas the poly(P) level was below the detectable level in cells that overproduced exopoly(P)ase. When exopoly(P)ase-overproducing cells were transformed again by a multiple copy number plasmid that carries the polyphosphate kinase gene (ppk), the cells accumulated a great amount of poly(P) and restored the UV and MMC sensitivities to the level of control cells. In exopoly(P)ase-overproducing cells, the expression of recA and umuDC were not induced by MMC. In addition, a strain containing multiple copies of ppk accumulated not only a large amount of poly(P) but also recA mRNA. Since recA expression was induced in a recA-deletion strain harboring a plasmid with the ppk gene, poly(P) could be necessary for regulating the expression of SOS genes without depending on the RecA-LexA regulatory network.  相似文献   

3.
脂蛋白(a) [ LP(a)]是一种与低密度脂蛋白(LDL)结构极其相似的脂蛋白,它由LDL脂质核心、载脂蛋白B100(apoB100)及特异性的成分载脂蛋白(a)[ apo(a)]组成. 大量的研究表明,高LP(a)是动脉粥样硬化独立的危险因素.而LP(a)在血浆中的水平及致病能力取决于其合成的速率及其颗粒的大小. 因此, 如何抑制LP(a)合成,进而从源头减少LP(a) 的血浆水平,对动脉粥样硬化的防治具有重要的意义.本文就当前关于影响LP(a)合成的环节及相关机制进行综述, 从而为降LP(a)药物的研究提供新的视角.  相似文献   

4.
Mock WL  Cheng H 《Biochemistry》2000,39(45):13945-13952
Hydroxamic acids of structure RCON(OH)CH(2)CH(CH(2)C(6)H(5))CO(2)H induce micromolar competitive inhibition of catalysis for the enzyme carboxypeptidase A. Enzyme affinity depends on the nature of the acyl group, for RCO equaling HCO, CH(3)CO, FCH(2)CO, F(2)CHCO, F(3)CCO, CH(3)OCH(2)CO, or CH(3)OCO. In acid dissociation these residues yield hydroxamic acid pK(a) values that vary from 7.6 to 10.3. Profiles of inhibitory pK(i) plotted versus pH indicate characteristically a maximum effectiveness near neutrality. Weaker binding to enzyme is generally displayed in either acidic or alkaline solution, with the position of the alkaline limb of the profiles depending on the pK(a) of the inhibitor. A reverse-protonation pattern of association with the enzyme is indicated, in which the hydroxamate anion of the inhibitor displaces a relatively acidic H(2)O ligand (pK(a) of 6) from the active-site zinc ion of carboxypeptidase A. The metal-coordinating, N-substituted hydroxamic acid functional groups exist in solution as a mixture of syn and anti rotamers, with relative abundances that depend on their pK(a). A pyrrolidinone analogue having a conformationally syn-fixed cyclohydroxamic acid was not an especially potent inhibitor. Structure-activity relationships suggest design criteria for hydroxamic acid inhibitors in order to provide most effective binding with metalloenzymes.  相似文献   

5.
In the past, the floristic diversity of arable fields has been described in terms of species diversity (SD) and their degree of coverage (C), but never in combination with the recording of the actually flowered species (FS) and their flowering intensity (FI) to striking differences in the cultivation methods on arable land. In relation to SD and C, however, FS and FI may provide important additional information on the functional biodiversity of fields. The aim was therefore to investigate the effects of (a) conventional, (b) organic, and (c) smallholder (never application of herbicides) on the floristic diversity. Using a region in Germany, we investigated SD, C, FS, and FI synchronously in (a), (b), and (c), by 356 vegetation surveys (5 × 5 m plots) conducted in spring and summer in 2019 in winter cereals. Statistical tests were used to analyze the differences between (a), (b), and (c). The medians were used to compare the floristic diversity of (a), (b), and (c) and finally relationships of FS and FI to SD were analyzed in relation to the cultivation methods. Significant differences in SD, C, FS, and FI were found between the (a), (b), and (c) in spring and summer characterized by sharp declines from (c) to (b) to (a). A drastic reduction in floristic diversity from (c) 100 to (b) 52 to (a) 3 was determined. Plants in flower (FS, FI) were very poorly in (a), moderately well to well in (b), and well to very well represented in (c). (C) to (a) was characterized by a sharp decline and from (a) to (b) by sharp increase in floristic diversity. With current acreage proportions of (a) in mind, this would affect, about one third of land area in Germany, associated with a drastic reduction in functional biodiversity for insects.  相似文献   

6.
Phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P(2)] plays a key role in the modulation of actin polymerization and vesicle trafficking. These processes seem to depend on the enrichment of PI(4,5)P(2) in plasma membrane domains. Here, we show that PI(4,5)P(2) does not form domains when in a fluid phosphatidylcholine matrix in the pH range of 4.8-8.4. This finding is at variance with the spontaneous segregation of PI(4,5)P(2) to domains as a mechanism for the compartmentalization of PI(4,5)P(2) in the plasma membrane. Water/bilayer partition of PI(4,5)P(2) is also shown to be dependent on the protonation state of the lipid.  相似文献   

7.
This study is concerned with validating the measurement of the plasma half-life of 11alpha-(2)H cortisol in an attempt to accurately assess the in vivo activity of 11beta-HSD2 in man. Oral administration of 5mg of cortisol-(13)C(4),(2)H(1) to a human subject after repeated ingestions of 130mg/day of glycyrrhetinic acid for 5 days resulted in a decrease in the rate constant of the cortisol-(13)C(4),(2)H(1) to cortisone-(13)C(4) conversion, a direct index reflecting 11beta-HSD2 activity. The reduced 11beta-HSD2 activity led to an increase in the elimination half-life of cortisol-(13)C(4),(2)H(1), indicating that the loss of 11alpha-(2)H is a sensitive in vivo means of assessing 11beta-HSD2 activity. A simultaneous oral administration of 3mg each of [1,2,4,19-(13)C(4),11alpha-(2)H]cortisol (cortisol-(13)C(4),(2)H(1)) and 11alpha-(2)H cortisol to another human subject confirmed the bioequivalency of the two labeled cortisols. The information obtained from the kinetic analysis of the 11beta-HSD2-catalyzed conversion of cortisol-(13)C(4),(2)H(1) to cortisone-(13)C(4) indicated that the elimination half-life of 11alpha-(2)H cortisol was a sensitive index of renal 11beta-HSD2 activity. The use of 11alpha-(2)H cortisol as a tracer appears to offer a significant advance in evaluating human 11beta-HSD2 activity in vivo.  相似文献   

8.
Peptidase mutants of Salmonella typhimurium   总被引:43,自引:30,他引:13       下载免费PDF全文
Six peptidase activities have been distinguished electrophoretically in cell extracts of Salmonella typhimurium with the aid of a histochemical stain. The activities can also be partially separated by chromatography on diethylaminoethyl-cellulose. These peptidases show overlapping substrate specificities. Mutants (pepN) of the parent strain leu-485 lacking one of these enzymes (peptidase N) were obtained by screening for colonies that do not hydrolyze the chromogenic substrate l-alanyl-beta-naphthylamide. The absence of this broad-specificity peptidase in leu-485 pepN(-) mutants allowed the selection of mutants unable to use l-leucyl-l-alaninamide as a leucine source. These mutants (leu-485 pepN(-)pepA(-)) lack a broad-specificity peptidase (peptidase A) similar to aminopeptidase I previously described in Escherichia coli. Mutants (pepD) lacking a dipeptidase (peptidase D) have been isolated from a leu-485 pepN(-)pepA(-) parent by penicillin selection for mutants unable to use l-leucyl-l-glycine as a leucine source. Mutants (pepB) lacking a fourth peptidase (peptidase B) have been isolated from a leu-485 pepN(-)pepA(-)pepD(-) strain by penicillin selection for failure to utilize l-leucyl-l-leucine as a source of leucine. Single recombinants were obtained by transduction for each of the peptidases missing in a leu-485 pepN(-)pepA(-)pepD(-)pepB(-) strain. The growth response of these recombinants to leucine peptides shows that all of these peptidases can function in the catabolism of peptides and that they display overlapping substrate specificities in vivo.  相似文献   

9.
Multidrug resistance protein 1 (MRP1) is capable of actively transporting a wide range of conjugated and unconjugated organic anions. The protein can also transport additional conjugated and unconjugated compounds in a GSH- or S-methyl GSH-stimulated manner. How MRP1 binds and transports such structurally diverse substrates is not known. We have used [(3)H]leukotriene C(4) (LTC(4)), a high affinity glutathione-conjugated physiological substrate, to photolabel intact MRP1, as well as fragments of the protein expressed in insect cells. These studies revealed that: (i) LTC(4) labels sites in the NH(2)- and COOH-proximal halves of MRP1, (ii) labeling of the NH(2)-half of MRP1 is localized to a region encompassing membrane-spanning domain (MSD) 2 and nucleotide binding domain (NBD) 1, (iii) labeling of this region is dependent on the presence of all or part of the cytoplasmic loop (CL3) linking MSD1 and MSD2, but not on the presence of MSD1, (iv) labeling of the NH(2)-proximal site is preferentially inhibited by S-methyl GSH, (v) labeling of the COOH-proximal half of the protein occurs in a region encompassing transmembrane helices 14-17 and appears not to require NBD2 or the cytoplasmic COOH-terminal region of the protein, (vi) labeling of intact MRP1 by LTC(4) is strongly attenuated in the presence of ATP and vanadate, and this decrease in labeling is attributable to a marked reduction in LTC(4) binding to the NH(2)-proximal site, and (vii) the attenuation of LTC(4) binding to the NH(2)-proximal site is a consequence of ATP hydrolysis and trapping of Vi-ADP exclusively at NBD2. These data suggest that MRP1-mediated transport involves a conformational change, driven by ATP hydrolysis at NBD2, that alters the affinity with which LTC(4) binds to one of two sites composed, at least in part, of elements in the NH(2)-proximal half of the protein.  相似文献   

10.
Kinetics of biofilm nitrification   总被引:5,自引:0,他引:5  
The reaction rates (r(NH(4) (+) ) and r(NO(2) (-) )) in the two-step nitrification reaction were measured in a fluidized-sand-bed biofilm reactor under a range of steady-state conditions with respect to bulk NH(4) (+), NO(2) (-), and O(2) concentrations. It was shown from theory and experiment that under low NH(4) (+) concentration conditions, if the O(2)/NH(4) (+) concentration ratio in the bulk liquid is less than the stoichiometric coefficient (3.4 mg/mg), then oxygen will be rate limiting. In all experiments r(NO(2) (-) ) decreased more than r(NH(4) (+) ) under low oxygen conditions. This resulted in high NO(2) (-) effluent concentrations under low residence time conditions. The influence of the oxygen penetration effects on the relative values of r(NH(4) (+) ) and r(NO(2) (-) ) was experimentally shown to be caused either by the Nitrobacter location in the inner biofilm regions or by a K(m) effect for oxygen. Theoretical support of these findings was provided by a differential diffusion-reaction model which was used to simulate the experimental results.  相似文献   

11.
Barley seedlings grown in the dark with 10 mm KNO(3) have low levels of nitrate reductase activity even though large amounts of No(3) (-) accumulate in the leaves. When the leaves are excised and transferred to the light, there is an increase in nitrate reductase activity both in the presence and absence of exogenous NO(3) (-). When the leaves are transferred to a glucose solution (0.05 m) but kept in the dark, induction of nitrate reductase activity occurs only when fresh NO(3) (-) is added to the system.In dark-grown leaves, there are small traces of NO(3) (-) in a "metabolic pool." Addition of glucose does not alter this distribution. Light, on the other hand, results in an appreciable accumulation of NO(3) (-) in the metabolic pool. There is a linear correlation between nitrate reductase activity and the size of the metabolic NO(3) (-) pool. Our results thus suggest that NO(3) (-) accumulates in a storage pool when seedlings are grown in continuous darkness. The transfer of this NO(3) (-) to an active metabolic pool is mediated by light but not by glucose. We believe that this transfer of NO(3) (-) leads to the induction of nitrate reductase. When NO(3) (-) is included in the medium, both light and glucose increase its incorporation into the metabolic pool. The results suggest two mechanisms for regulating the metabolic NO(3) (-) pool: (a) a transfer from the storage pool which requires light; and (b) a transfer from the external medium which requires either glucose or light.  相似文献   

12.
Electrostatic binding sites of extracellular polymeric substances (EPS) were characterized from titration data using linear programming analysis. Test results for three synthetic solutions of given solutes comprising amino, carboxyl, and phenolic groups indicated that this method was able to identify the electrostatic binding sites. For the six sites with pK(a) between 3 and 10, the estimated pK(a) deviated 0.11 +/- 0.09 from the theoretical values, and the estimated concentrations deviated 3.0% +/- 0.9% from the actual concentrations. Two EPS samples were then extracted from a hydrogen-producing sludge (HPS) and a sulfate-reducing biofilm (SRB). Analysis of charge excess data in titration from pH 3 to 11 indicated that the EPS of HPS comprised of five electrostatic binding sites with pK(a) ranging from 3 to 11. The pK(a) values of these binding sites and the possible corresponding functional groups were pK(a) 4.8 (carboxyl), pK(a) 6.0 (carboxyl/phosphoric), pK(a) 7.0 (phosphoric), pK(a) 9.8 (amine/phenolic), and pK(a) 11.0 (hydroxyl). EPS of the SRB comprised five of similar binding sites (with corresponding pK(a) values of 4.4, 6.0, 7.4, 9.4, and 11.0), plus one extra site at pK(a) 8.2, which was likely corresponding to the sulfhydryl group. The total electrostatic binding site concentration of EPS extracted from HPS were 10.88 mmol/g-EPS, of which the highest concentration was from the site of pK(a) 11.0. The corresponding values for the EPS extracted from SRB were 16.44 mmol/g-EPS and pK(a) 4.4. The total concentrations of electrostatic binding sites found in this study were 20- to 30-fold of those reported for bacterial cell surface, implying that EPS might be more crucial in biosorption of metals than bacterial cell surface in wastewater treatment and in bioremediation.  相似文献   

13.
Mechanism of reaction of myeloperoxidase with nitrite   总被引:10,自引:0,他引:10  
Myeloperoxidase (MPO) is a major neutrophil protein and may be involved in the nitration of tyrosine residues observed in a wide range of inflammatory diseases that involve neutrophils and macrophage activation. In order to clarify if nitrite could be a physiological substrate of myeloperoxidase, we investigated the reactions of the ferric enzyme and its redox intermediates, compound I and compound II, with nitrite under pre-steady state conditions by using sequential mixing stopped-flow analysis in the pH range 4-8. At 15 degrees C the rate of formation of the low spin MPO-nitrite complex is (2.5 +/- 0.2) x 10(4) m(-1) s(-1) at pH 7 and (2.2 +/- 0.7) x 10(6) m(-1) s(-1) at pH 5. The dissociation constant of nitrite bound to the native enzyme is 2.3 +/- 0.1 mm at pH 7 and 31.3 +/- 0.5 micrometer at pH 5. Nitrite is oxidized by two one-electron steps in the MPO peroxidase cycle. The second-order rate constant of reduction of compound I to compound II at 15 degrees C is (2.0 +/- 0.2) x 10(6) m(-1) s(-1) at pH 7 and (1.1 +/- 0.2) x 10(7) m(-1) s(-1) at pH 5. The rate constant of reduction of compound II to the ferric native enzyme at 15 degrees C is (5.5 +/- 0.1) x 10(2) m(-1) s(-1) at pH 7 and (8.9 +/- 1.6) x 10(4) m(-1) s(-1) at pH 5. pH dependence studies suggest that both complex formation between the ferric enzyme and nitrite and nitrite oxidation by compounds I and II are controlled by a residue with a pK(a) of (4.3 +/- 0.3). Protonation of this group (which is most likely the distal histidine) is necessary for optimum nitrite binding and oxidation.  相似文献   

14.
The inositol lipid and phosphate binding properties and the cellular localization of phospholipase Cdelta(4) (PLCdelta(4)) and its isolated pleckstrin homology (PH) domain were analyzed in comparison with the similar features of the PLCdelta(1) protein. The isolated PH domains of both proteins showed plasma membrane localization when expressed in the form of a green fluorescent protein fusion construct in various cells, although a significantly lower proportion of the PLCdelta(4) PH domain was membrane-bound than in the case of PLCdelta(1)PH-GFP. Both PH domains selectively recognized phosphatidylinositol 4,5-bisphosphate (PI(4,5)P(2)), but a lower binding of PLCdelta(4)PH to lipid vesicles containing PI(4,5)P(2) was observed. Also, higher concentrations of inositol 1,4,5-trisphosphate (Ins(1,4,5)P(3)) were required to displace the PLCdelta(4)PH from the lipid vesicles, and a lower Ins(1,4,5)P(3) affinity of PLCdelta(4)PH was found in direct Ins(1,4,5)P(3) binding assays. In sharp contrast to the localization of its PH domain, the full-length PLCdelta(4) protein localized primarily to intracellular membranes mostly to the endoplasmic reticulum (ER). This ER localization was in striking contrast to the well documented PH domain-dependent plasma membrane localization of PLCdelta(1). A truncated PLCdelta(4) protein lacking the entire PH domain still showed the same ER localization as the full-length protein, indicating that the PH domain is not a critical determinant of the localization of this protein. Most important, the full-length PLCdelta(4) enzyme still showed binding to PI(4,5)P(2)-containing micelles, but Ins(1,4,5)P(3) was significantly less potent in displacing the enzyme from the lipid than with the PLCdelta(1) protein. These data suggest that although structurally related, PLCdelta(1) and PLCdelta(4) are probably differentially regulated in distinct cellular compartments by PI(4,5)P(2) and that the PH domain of PLCdelta(4) does not act as a localization signal.  相似文献   

15.
A series of 12 bovine pancreatic trypsin inhibitor variants mutated in the P(4) and P(3) positions of the canonical binding loop containing additional K15R and M52L mutations were used to probe the role of single amino acid substitutions on binding to bovine trypsin and to the following human proteinases involved in blood clotting: plasmin, plasma kallikrein, factors X(a) and XII(a), thrombin, and protein C. The mutants were expressed in Escherichia coli as fusion proteins with the LE1413 hydrophobic polypeptide and purified from inclusion bodies; these steps were followed by CNBr cleavage and oxidative refolding. The mutants inhibited the blood-clotting proteinases with association constants in the range of 10(3)-10(10) m(-)(1). Inhibition of plasma kallikrein, factors X(a) and XII(a), thrombin, and protein C could be improved by up to 2 orders of magnitude by the K15R substitution. The highest increase in the association constant for P(3) mutant was measured for factor XII(a); P13S substitution increased the K(a) value 58-fold. Several other substitutions at P(3) resulted in about 10-fold increase for factor X(a), thrombin, and protein C. The cumulative P(3) and P(1) effects on K(a) values for the strongest mutant compared with the wild type bovine pancreatic trypsin inhibitor were in the range of 2.2- (plasmin) to 4,000-fold (factors XII(a) and X(a)). The substitutions at the P(4) site always caused negative effects (a decrease in the range from over 1,000- to 1.3-fold) on binding to all studied enzymes, including trypsin. Thermal stability studies showed a very large decrease of the denaturation temperature (about 22 degrees C) for all P(4) mutants, suggesting that substitution of the wild type Gly-12 residue leads to a change in the binding loop conformation manifesting itself in non-optimal binding to the proteinase active site.  相似文献   

16.
Thermus aquaticus contains four major fatty acids, iso-C(15) (28%), iso-C(16) (9%), normal-C(16) (13%), and iso-C(17) (48%), when grown at 70 C, as determined by gas chromatography and mass spectrometry. Small amounts of iso-C(12), normal-C(12:1), iso-C(13), normal-C(14), iso-C(14), and normal-C(15:1) were also detected. A change in growth temperature (50 to 75 C at 5-C intervals) affects a shift in the proportions of some of the fatty acids. The proportions of the monoenoic and branched-C(17) fatty acids decreased and the proportions of the higher-melting iso-C(16) and normal-C(16) fatty acids increased. Cells grown at 75 C contained 70% more total fatty acids than cells grown at 50 C. The largest increases, in absolute amounts, were in the content of iso-C(16) and normal-C(16) fatty acids, with only a 1.6-fold increase in the major iso-C(15) and iso-C(17) fatty acids. There was a 2.5-fold decrease in normal-C(15:1) and at least a 24-fold decrease in anteiso-C(17), which is present at 50 and 55 C but not at higher temperatures. There was no difference in proportion or amount of fatty acids between exponential and stationary-phase cells grown at 70 C. When cells were grown on glutamate instead of yeast-extract and tryptone at 70 C, the total fatty acid content remained constant, but there was an increase in the proportions of iso-C(16) and normal-C(16) fatty acids concomitant with a decrease in the proportions of the iso-C(15) and iso-C(17) fatty acids.  相似文献   

17.
The effect of alpha-tocopherol (alpha-tp) prepared in solvents of different polarity in a wide range of concentrations (10(-4) M - 10(-25) M) on lipid phase structural characteristics of microsomal membranes isolated from mouse liver cells has been investigated in vitro. Structural changes in membranes were detected on a Bruker-200D ESR-spectrometer (Germany) by the method of spin probes. Changes in the rigidity of surface lipid bilayer regions (8 A) and microviscosity of deep membrane layers (20 A) were studied using the stable nitroxyl radicals 5- and 16-doxylstearic acids, correspondingly. As a result, nonlinear multimodal dose dependences were obtained. It was demonstrated that the physiological (10(-4) M - 10(-9) M) and ultralow doses of alpha-tocopherol up to "apparent" concentrations (10(-11) M - 10(-25) M) increased the rigidity of surface lipid bilayer regions and microviscosity in the depth of membrane. Additionally, these doses of alpha-tp induced an increase in the number of thermoinduced structural transitions in deep lipid bilayer regions. The effect at "apparent" concentrations (< 10(-18) M) has only been observed in polar alpha-tocopherol solutions. The results obtained are statistically reliable with a significance level of 95%.  相似文献   

18.
Yamazaki S  Morioka C  Itoh S 《Biochemistry》2004,43(36):11546-11553
Tyrosinase is a copper monooxygenase containing a coupled dinuclear copper active site (type-3 copper), which catalyzes oxygenation of phenols (phenolase activity) as well as dehydrogenation of catechols (catecholase activity) using O(2) as the oxidant. In this study, catalase activity (conversion of H(2)O(2) to (1/2)O(2) and H(2)O) and peroxygenase activity (H(2)O(2)-dependent oxygenation of substrates) of mushroom tyrosinase have been examined kinetically by using amperometric O(2) and H(2)O(2) sensors. The catalase activity has been examined by monitoring the initial rate of O(2) production from H(2)O(2) in the presence of a catalytic amount of tyrosinase in 0.1 M phosphate buffer (pH 7.0) at 25 degrees C under initially anaerobic conditions. It has been found that the catalase activity of mushroom tyrosinase is three-order of magnitude greater than that of mollusk hemocyanin. The higher catalase activity of tyrosinase could be attributed to easier accessibility of H(2)O(2) to the dinuclear copper site of tyrosinase. Mushroom tyrosinase has also been demonstrated for the first time to catalyze oxygenation reaction of phenols with H(2)O(2) (peroxygenase activity). The reaction has been investigated kinetically by monitoring the H(2)O(2) consumption rate in 0.5 M borate buffer (pH 7.0) under aerobic conditions. Similarity of the substituent effects of a series of p-substituted phenols in the peroxygenase reaction with H(2)O(2) to those in the phenolase reaction with O(2) as well as the absence of kinetic deuterium isotope effect with a perdeuterated substrate (p-Cl-C(6)D(4)OH vs p-Cl-C(6)H(4)OH) clearly demonstrated that the oxygenation mechanisms of phenols in both systems are the same, that is, the electrophilic aromatic substitution reaction by a (micro-eta(2):eta(2)-peroxo)dicopper(II) intermediate of oxy-tyrosinase.  相似文献   

19.
20.
Two strains of minute virus of mice (MVM) show different host cell specificities. The prototype strain MVM(p) grows in fibroblasts, whereas the immunosuppressive variant MVM(i) grows in T lymphocytes. In this study, we have mapped on the viral genome a cell type-specific determinant: it is located between 69 and 85 map units in a region coding for the viral capsid proteins. The DNA of MVM(p) does not replicate in lymphocytes. MVM(i) cannot help MVM(p) grow in lymphocytes; thus the determinant acts in a cis fashion. We did not detect viral mRNA during a restrictive infection of lymphocytes with MVM(p). However, when the same cells were transfected with cloned DNA, both MVM(p) and MVM(i) DNAs were transcribed with the same efficiency from both promoters and the RNA was processed normally. Therefore, the specificity determinant is not a cell type-specific enhancer.  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