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.
Isolated copper/zinc superoxide dismutase (Cu/Zn-SOD) or manganese superoxide dismutase (Mn-SOD) together with hydrogen peroxide (H(2)O(2)) caused rapid breakdown of nitric oxide (NO) and production of peroxynitrite (ONOO(-)) indicated by the oxidation of dihydrorhodamine-1,2,3 (DHR) to rhodamine-1,2,3. The breakdown of NO by this reaction was inhibited by cyanide (CN(-)) or by diethyldithiocarbamate (DETC), both Cu/Zn-SOD inhibitors, and the conversion of DHR to rhodamine-1,2,3 was inhibited by incubating Cu/Zn-SOD with either CN(-) or with high levels of H(2)O(2) or by including urate, a potent scavenger of ONOO(-). In the presence of phenol, the reaction of SOD, H(2)O(2) and NO caused nitration of phenol, which is known to be a footprint of ONOO(-) formation. H(2)O(2) addition to macrophages (cell line J774) expressing the inducible form of NO synthase (i-NOS) caused rapid breakdown of the NO they produced and this was also inhibited by CN(-) and by DETC. Subsequent ONOO(-) production by the macrophages, via this reaction, was inhibited by CN(-), high levels of H(2)O(2) or by urate. H(2)O(2) addition to i-NOS macrophages also caused cell death which was, in part, prevented by DETC or urate. We also found inhibition of mitochondrial respiration with malate and pyruvate as substrates, when isolated liver mitochondria were incubated with Cu/Zn-SOD, H(2)O(2) and NO. Inhibition of mitochondrial respiration was partly prevented by urate. The production of ONOO(-) by SOD may be of significant importance pathologically under conditions of elevated H(2)O(2) and NO levels, and might contribute to cell death in inflammatory and neurodegenerative diseases, as well as in macrophage-mediated host defence.  相似文献   

2.
Aflatoxin B(1) (AFB(1) ) is a potent mycotoxin with mutagenic, carcinogenic, teratogenic, hepatotoxic, and immunosuppressive properties. In order to develop a bioremediation system for AFB(1) -contaminated foods by white-rot fungi or ligninolytic enzymes, AFB(1) was treated with manganese peroxidase (MnP) from the white-rot fungus Phanerochaete sordida YK-624. AFB(1) was eliminated by MnP. The maximum elimination (86.0%) of AFB(1) was observed after 48 h in a reaction mixture containing 5 nkat of MnP. The addition of Tween 80 enhanced AFB(1) elimination. The elimination of AFB(1) by MnP considerably reduced its mutagenic activity in an umu test, and the treatment of AFB(1) by 20 nkat MnP reduced the mutagenic activity by 69.2%. (1) H-NMR and HR-ESI-MS analysis suggested that AFB(1) is first oxidized to AFB(1) -8,9-epoxide by MnP and then hydrolyzed to AFB(1) -8,9-dihydrodiol. This is the first report that MnP can effectively remove the mutagenic activity of AFB(1) by converting it into AFB(1) -8,9-dihydrodiol.  相似文献   

3.
Activation of ERK-1 and -2 by H(2)O(2) in a variety of cell types requires epidermal growth factor receptor (EGFR) phosphorylation. In this study, we investigated the activation of ERK by ONOO(-) in cultured rat lung myofibroblasts. Western blot analysis using anti-phospho-ERK antibodies along with an ERK kinase assay using the phosphorylated heat- and acid-stable protein (PHAS-1) substrate demonstrated that ERK activation peaked within 15 min after ONOO(-) treatment and was maximally activated with 100 micrometer ONOO(-). Activation of ERK by ONOO(-) and H(2)O(2) was blocked by the antioxidant N-acetyl-l-cysteine. Catalase blocked ERK activation by H(2)O(2), but not by ONOO(-), demonstrating that the effect of ONOO(-) was not due to the generation of H(2)O(2). Both H(2)O(2) and ONOO(-) induced phosphorylation of EGFR in Western blot experiments using an anti-phospho-EGFR antibody. However, the EGFR tyrosine kinase inhibitor AG1478 abolished ERK activation by H(2)O(2), but not by ONOO(-). Both H(2)O(2) and ONOO(-) activated Raf-1. However, the Raf inhibitor forskolin blocked ERK activation by H(2)O(2), but not by ONOO(-). The MEK inhibitor PD98059 inhibited ERK activation by both H(2)O(2) and ONOO(-). Moreover, ONOO(-) or H(2)O(2) caused a cytotoxic response of myofibroblasts that was prevented by preincubation with PD98059. In a cell-free kinase assay, ONOO(-) (but not H(2)O(2)) induced autophosphorylation and nitration of a glutathione S-transferase-MEK-1 fusion protein. Collectively, these data indicate that ONOO(-) activates EGFR and Raf-1, but these signaling intermediates are not required for ONOO(-)-induced ERK activation. However, MEK-1 activation is required for ONOO(-)-induced ERK activation in myofibroblasts. In contrast, H(2)O(2)-induced ERK activation is dependent on EGFR activation, which then leads to downstream Raf-1 and MEK-1 activation.  相似文献   

4.
Lower esophageal sphincter (LES) tone depends on PGF(2alpha) and thromboxane A(2) acting on receptors linked to G(i3) and G(q) to activate phospholipases and produce second messengers resulting in muscle contraction. We therefore examined PGF(2alpha) signal transduction in circular smooth muscle cells isolated by enzymatic digestion from cat esophagus (Eso) and LES. In Eso, PGF(2alpha)-induced contraction was inhibited by antibodies against the alpha-subunit of G(13) and the monomeric G proteins RhoA and ADP-ribosylation factor (ARF)1 and by the C3 exoenzyme of Clostridium botulinum. A [(35)S]GTPgammaS-binding assay confirmed that G(13), RhoA, and ARF1 were activated by PGF(2alpha). Contraction of Eso was reduced by propranolol, a phospholipase D (PLD) pathway inhibitor and by chelerythrine, a PKC inhibitor. In LES, PGF(2alpha)-induced contraction was inhibited by antibodies against the alpha-subunit of G(q) and G(i3), and a [(35)S]GTPgammaS-binding assay confirmed that G(q) and G(i3) were activated by PGF(2alpha). PGF(2alpha)-induced contraction of LES was reduced by U-73122 and D609 and unaffected by propranolol. At low PGF(2alpha) concentration, contraction was blocked by chelerythrine, whereas at high concentration, contraction was blocked by chelerythrine and CGS9343B. Thus, in Eso, PGF(2alpha) activates a PLD- and protein kinase C (PKC)-dependent pathway through G(13), RhoA, and ARF1. In LES, PGF(2alpha) receptors are coupled to G(q) and G(i3), activating phosphatidylinositol- and phosphatidylcholine-specific phospholipase C. At low concentrations, PGF(2alpha) activates PKC. At high concentration, it activates both a PKC- and a calmodulin-dependent pathway.  相似文献   

5.
1,25-Dihydroxyvitamin D(3) (1,25(OH)(2)D(3)) treatment of osteoblastic cells was shown previously to attenuate Parathyroid hormone (PTH) response by inhibiting adenylyl cyclase (AC) activity. In this study, we have investigated the mechanism by which 1,25(OH)(2)D(3) inhibits AC in rat osteoblastic UMR 106-01 cells. 1,25(OH)(2)D(3) treatment inhibited both PTH and forskolin-stimulated AC activity by 25%-50% within 12 min in a concentration-dependent manner suggesting a direct inhibition of the AC enzyme. Treatment with 25(OH)D(3) had no effect on basal or stimulated AC activity. We determined the profile of AC subtypes expressed in UMR cells and found AC VI to be the dominant subtype accounting for 50% of AC mRNA. Since AC VI can be inhibited by protein kinase C (PKC) phosphorylation, we examined 1,25(OH)(2)D(3) activation of various PKC isoforms. 1,25(OH)(2)D(3) increased the membrane translocation of PKC-betaI, -delta, and -zeta with a concomitant increase in PKC activity. The translocation of PKC-betaI and -delta was blocked by the PLC inhibitor U73122 whereas that of PKC-zeta was abolished by the PI-3 kinase inhibitor wortmannin. The attenuation of cAMP production by 1,25(OH)(2)D(3) was antagonized by the PKC inhibitors Go6850, calphostin C, and wortmannin, but not by a calmodulin kinase II (CaMKII) inhibitor. Treatment with 1,25(OH)(2)D(3) for 20 min increased AC VI phosphorylation by 10.8-fold and this was blocked partially by Go6850 and partially by wortmannin but was unaffected by CaMKII inhibitor. These results demonstrate that 1,25(OH)(2)D(3) activation of PKC isoforms leads to phosphorylation of AC VI and inhibition of PTH-activation of this pathway in osteoblasts.  相似文献   

6.
The effects of ribo- and deoxyribonucleic acids on the activity of detergent-dispersed adenylate cyclases from rat and bovine brain were examined. Mn2+ (10 mM)-activated adenylate cyclase was inhibited by micromolar concentrations of poly(A) (IC50 congruent to 0.45 microM). This inhibition was directly due to poly(A) and was not mediated by: (a) protein contamination of the poly(A) preparation, (b) metal chelation, (c) formation of an acid-soluble inhibitor of adenylate cyclase, (d) effects on the specific activity of [alpha-32P]ATP, (e) competition with MnATP for binding to adenylate cyclase, or (f) diversion of substrate to an alternate polymerase reaction. Inhibition of adenylate cyclase by poly(A) was on the enzyme's catalytic unit, as purified preparations of the enzyme from bovine brain were inhibited by poly(A). This inhibition by poly(A) was not likely mediated via the enzyme's "P"-site, through which activated forms of the enzyme are selectively inhibited by specific adenosine phosphates. In contrast with inhibition by the "P"-site agonist 3' AMP, inhibition of adenylate cyclase by poly(A) was slow in onset and was not reversible by dilution and showed a different metal-dependence. Inhibition of adenylate cyclase was relatively specific for poly(A) as poly(U) caused less than 50% inhibition and deoxyribonucleic acids had no effect. The potency and specificity of the inhibition of adenylate cyclase by poly(A) imply a biochemically interesting interaction that is possibly also of physiological significance.  相似文献   

7.
Prostaglandin (PG)E derivatives are widely used for treating gastric mucosal injury. PGE receptors are classified into four subtypes, EP(1), EP(2), EP(3), and EP(4). We have tested which EP receptor subtypes participate in gastric mucosal protection against ethanol-induced gastric mucosal injury and clarified the mechanisms of such protection. The gastric mucosa of anesthetized rats was perfused at 2 ml/min with physiological saline, agonists for EP(1), EP(2), EP(3), and EP(4), or 50% ethanol, using a constant-rate pump connected to a cannula placed in the esophagus. The gastric microcirculation of the mucosal base of anesthetized rats was observed by transillumination through a window made by removal of the adventitia and muscularis externa. PGE(2) and subtype-specific EP agonists were applied to the muscularis mucosae at the window. Application of 50% ethanol dilated the mucosal arterioles and constricted the collecting venules. Collecting venule constriction by ethanol was completely inhibited by PGE(2) and by EP(2) and EP(4) agonists (100 nM) but not by an EP(1) or an EP(3) agonist. Ethanol-induced mucosal injury was also inhibited by EP(2) and EP(4) agonists. When leukotriene (LT)C(4) levels in the perfusate of the gastric mucosa were determined by ELISA, intragastric ethanol administration elevated the LTC(4) levels sixfold from the basal levels. These elevated levels were significantly (60%) reduced by both EP(2) and EP(4) agonists but not by other EP agonists. Since LTC(4) application at the window constricted collecting venules strongly, and an LTC antagonist reduced ethanol-induced mucosal injury, reductions in LTC(4) generation in response to EP(2) and EP(4) receptor signaling may be relevant to the protective action of PGE(2). The present results indicate that EP(2) and EP(4) receptor signaling inhibits ethanol-induced gastric mucosal injury through cancellation of collecting venule constriction by reducing LTC(4) production.  相似文献   

8.
The aim of this work was to characterize the placental uptake of folic acid from the maternal circulation. Using 2 human trophoblast cell lines (BeWo and JAR), we verified that uptake of 3H-folic acid was pH-dependent, increasing significantly with decreasing extracellular pH. In BeWo cells, uptake of 3H-folic acid at pH 5.5 was (i) Na+-independent; (ii) inhibited by folic acid, 5-methyltetrahydrofolate (5-MTHF), and methotrexate (MTX); (iii) inhibited by the anion transport inhibitors 4,4'-diisothiocyanatostilbene-2,2'-disulphonic acid (DIDS) and 4-acetamido-4'-isothiocyano-2,2'-disulfonic acid stilbene (SITS); (iv) inhibited by the proton ionophore carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone (FCCP); (v) not inhibited by blockers of receptor-mediated endocytosis (cytochalasin D and monensin); (vi) trans-inhibited by MTX and folic acid; and (vii) not affected by an anti-reduced folate transporter-1 (RFC) antibody. At pH 7.5, uptake of 3H-folic acid was (i) Na+-independent; (ii) inhibited by folic acid and MTX, but not by 5-MTHF; (iii) inhibited by SITS, but not by DIDS; (iv) not affected by FCCP; (v) inhibited by monensin (but not by cytochalasin D); (vi) trans-inhibited by folic acid (but not by MTX); and (vii) inhibited by an anti-RFC antibody. In conclusion, in BeWo cells, both RFC and receptor-mediated endocytosis seem to be involved in 3H-folic acid uptake at pH 7.5, whereas at pH 5.5, RFC and (or) a low pH-operating transporter distinct from RFC are involved.  相似文献   

9.
Stimulation of phospholipase C (PLC) by G(q)-coupled receptors such as the M(3) muscarinic acetylcholine receptor (mAChR) is caused by direct activation of PLC-beta enzymes by Galpha(q) proteins. We have recently shown that G(s)-coupled receptors can stimulate PLC-epsilon, apparently via formation of cyclic AMP and activation of the Ras-related GTPase Rap2B. Here we report that PLC stimulation by the M(3) mAChR expressed in HEK-293 cells also involves, in part, similar mechanisms. M(3) mAChR-mediated PLC stimulation and [Ca(2+)](i) increase were reduced by 2',5'-dideoxyadenosine (dd-Ado), a direct adenylyl cyclase inhibitor. On the other hand, overexpression of Galpha(s) or Epac1, a cyclic AMP-regulated guanine nucleotide exchange factor for Rap GTPases, enhanced M(3) mAChR-mediated PLC stimulation. Inactivation of Ras-related GTPases with clostridial toxins suppressed the M(3) mAChR responses. The inhibitory toxin effects were mimicked by expression of inactive Rap2B, but not of other inactive GTPases (Rac1, Ras, RalA, Rap1A, and Rap2A). Activation of the M(3) mAChR induced GTP loading of Rap2B, an effect strongly enhanced by overexpression of Galpha(s) and inhibited by dd-Ado. Overexpression of PLC-epsilon and PLC-beta1, but not PLC-gamma1 or PLC-delta1, enhanced M(3) mAChR-mediated PLC stimulation and [Ca(2+)](i) increase. In contrast, expression of a catalytically inactive PLC-epsilon mutant reduced PLC stimulation by the M(3) mAChR and abrogated the potentiating effect of Galpha(s). In conclusion, our findings suggest that PLC stimulation by the M(3) mAChR is a composite action of PLC-beta1 stimulation by Galpha(q) and stimulation of PLC-epsilon apparently mediated by G(s)-dependent cyclic AMP formation and subsequent activation of Rap2B.  相似文献   

10.
Bovine pancreatic ribonuclease A (RNase A) catalyzes the cleavage of the P-O(5') bond in RNA. Although this enzyme has been the object of much landmark work in bioorganic chemistry, the nature of its rate-limiting transition state and its catalytic rate enhancement had been unknown. Here, the value of k(cat)/K(m) for the cleavage of UpA by wild-type RNase A was found to be inversely related to the concentration of added glycerol. In contrast, the values of k(cat)/K(m) for the cleavage of UpA by a sluggish mutant of RNase A and the cleavage of the poor substrate UpOC(6)H(4)-p-NO(2) by wild-type RNase A were found to be independent of glycerol concentration. Yet, UpA cleavage by the wild-type and mutant enzymes was found to have the same dependence on sucrose concentration, indicating that catalysis of UpA cleavage by RNase A is limited by desolvation. The rate of UpA cleavage by RNase A is maximal at pH 6.0, where k(cat) = 1.4 × 10(3) s(-1) and k(cat)/K(m) = 2.3 × 10(6) M(-1)s(-1) at 25°C. At pH 6.0 and 25°C, the uncatalyzed rate of [5,6-(3)H]Up[3,5,8-(3)H]A cleavage was found to be k(uncat) = 5 × 10(-9) s(-1) (t(1/2) = 4 years). Thus, RNase A enhances the rate of UpA cleavage by 3 × 10(11)-fold by binding to the transition state for P-O(5') bond cleavage with a dissociation constant of <2 × 10(-15) M.  相似文献   

11.
Norepinephrine (NE) stimulates phospholipase D (PLD) through a Ras/MAPK pathway in rabbit vascular smooth muscle cells (VSMC). NE also activates calcium influx and calmodulin (CaM)-dependent protein kinase II-dependent cytosolic phospholipase A(2) (cPLA(2)). Arachidonic acid (AA) released by cPLA(2)-catalyzed phospholipid hydrolysis is then metabolized into hydroxyeicosatetraenoic acids (HETEs) through lipoxygenase and cytochrome P450 4A (CYP4A) pathways. HETEs, in turn, have been shown to stimulate Ras translocation and to increase MAPK activity in VSMC. This study was conducted to determine the contribution of cPLA(2)-derived AA and its metabolites (HETEs) to the activation of PLD. NE-induced PLD activation was reduced by two structurally distinct CaM antagonists, W-7 and calmidazolium, and by CaM-dependent protein kinase II inhibition. Blockade of cPLA(2) activity or protein depletion with selective cPLA(2) antisense oligonucleotides abolished NE-induced PLD activation. The increase in PLD activity elicited by NE was also blocked by inhibitors of lipoxygenases (baicalein) and CYP4A (17-octadecynoic acid), but not of cyclooxygenase (indomethacin). AA and its metabolites (12(S)-, 15(S)-, and 20-HETEs) increased PLD activity. PLD activation by AA and HETEs was reduced by inhibitors of Ras farnesyltransferase (farnesyl protein transferase III and BMS-191563) and MEK (U0126 and PD98059). These data suggest that HETEs are the mediators of cPLA(2)-dependent PLD activation by NE in VSMC. In addition to cPLA(2), PLD was also found to contribute to AA release for prostacyclin production via the phosphatidate phosphohydrolase/diacylglycerol lipase pathway. Finally, a catalytically inactive PLD(2) (but not PLD(1)) mutant inhibited NE-induced PLD activity, and PLD(2) was tyrosine-phosphorylated in response to NE by a MAPK-dependent pathway. We conclude that NE stimulates cPLA(2)-dependent PLD(2) through lipoxygenase- and CYP4A-derived HETEs via the Ras/ERK pathway by a mechanism involving tyrosine phosphorylation of PLD(2) in rabbit VSMC.  相似文献   

12.
The objective of this study was to get a more understandable picture of the mechanism underlying the anticonvulsant action of vinpocetine. The question of how the cerebral excitability is affected was investigated by determining the effect of vinpocetine on the changes on the internal concentrations of Na(+) (Na(i)) and Ca(2+) (Ca(i)) induced by different concentrations of the convulsing agent 4-aminopyridine (4-AP) in striatal isolated nerve endings. The cytosolic concentrations of Na(i) and Ca(i) were detected fluorimetrically with sodium-binding benzofuran isophthalate (SBFI) and fura-2, respectively. Vinpocetine, like the Na(+) channel blocker, tetrodotoxin, abolished the increase in Na(i) induced by 0.1 mM 4-AP and only inhibited in 30% the rise in Na(i) induced by 1mM 4-AP. In contrast with the different sensitivity of the rise in Na(i) induced by 0.1 and 1mM 4-AP to vinpocetine and tetrodotoxin, the rise in Ca(i) induced by the two concentrations of 4-AP was markedly inhibited by vinpocetine (and tetrodotoxin), indicating that only the voltage-sensitive sodium channels (VSSC)-mediated fraction of the rise in Na(i) induced by 4-AP is linked with the activation of pre-synaptic Ca(2+) channels. The elevation of Ca(2+) induced by high K(+) (30 mM) does not require a Na(+) gradient and is vinpocetine and tetrodotoxin insensitive. In contrast, the elevation of Ca(i) induced by 4-AP, requires a physiological (out/in) Na(+) gradient and is vinpocetine and tetrodotoxin-sensitive. It is concluded that by blocking the tetrodotoxin-sensitive fraction of the rise in Na(i) induced by 4-AP, vinpocetine inhibits the concomitant rise in Ca(i) induced by 4-AP. The inhibitory effect of vinpocetine on pre-synaptic voltage-sensitive sodium channels may underlie the in vivo anticonvulsant action of vinpocetine.  相似文献   

13.
In rat uterine stromal cells (U(III) cells), an oxidative stress induced by H(2)O(2) caused a dose-dependent release of arachidonic acid (AA) that was independent of intracellular Ca(2+) concentration and was not inhibited by Ca(2+)-dependent phospholipase A(2) (cPLA(2)) inhibitors, nor by protein kinase C (PKC) inhibitors or by PKC down-regulation. H(2)O(2) treatment did not impair AA esterification but significantly increased Ca(2+)-independent PLA(2) (iPLA(2)) activity. Since iPLA(2) specific inhibitor bromoenollactone almost completely suppressed the release of AA induced by H(2)O(2), we conclude that iPLA(2) activity represents the major mechanism by which H(2)O(2) increases the availability of non-esterified AA in U(III) cells. Moreover, PKC inhibitors sphingosine and calphostin C markedly potentiated the release of AA trigger by H(2)O(2), suggesting a regulatory mechanism of iPLA(2) by PKC that remains to be clarified.  相似文献   

14.
The role of H(2)O(2) and protein thiol oxidation in oxidative stress-induced epithelial paracellular permeability was investigated in Caco-2 cell monolayers. Treatment with a H(2)O(2) generating system (xanthine oxidase + xanthine) or H(2)O(2) (20 microM) increased the paracellular permeability. Xanthine oxidase-induced permeability was potentiated by superoxide dismutase and prevented by catalase. H(2)O(2)-induced permeability was prevented by ferrous sulfate and potentiated by deferoxamine and 1,10-phenanthroline. GSH, N-acetyl-L-cysteine, dithiothreitol, mercaptosuccinate, and diethylmaleate inhibited H(2)O(2)-induced permeability, but it was potentiated by 1,3-bis(2-chloroethyl)-1-nitrosourea. H(2)O(2) reduced cellular GSH and protein thiols and increased GSSG. H(2)O(2)-mediated reduction of GSH-to-GSSG ratio was prevented by ferrous sulfate, GSH, N-acetyl-L-cysteine, diethylmaleate, and mercaptosuccinate and potentiated by 1,10-phenanthroline and 1, 3-bis(2-chloroethyl)-1-nitrosourea. Incubation of soluble fraction of cells with GSSG reduced protein tyrosine phosphatase (PTPase) activity, which was prevented by coincubation with GSH. PTPase activity was also lower in H(2)O(2)-treated cells. This study indicates that H(2)O(2), but not O(2)(-). or.OH, increases paracellular permeability of Caco-2 cell monolayer by a mechanism that involves oxidation of GSH and inhibition of PTPases.  相似文献   

15.
Nitrogen and oxygen transformations were studied in a bioturbated (reworked by animals) estuarine sediment (Norsminde Fjord, Denmark) by using a combination of N isotope (NO(3)), specific inhibitor (C(2)H(2)), and microsensor (N(2)O and O(2)) techniques in a continuous-flow core system. The estuarine water was NO(3) rich (125 to 600 muM), and NO(3) was consistently taken up by the sediment on the four occasions studied. Total NO(3) uptake (3.6 to 34.0 mmol of N m day) corresponded closely to N(2) production (denitrification) during the experimental steady state, which indicated that dissimilatory, as well as assimilatory, NO(3) reduction to NH(4) was insignificant. When C(2)H(2) was applied in the flow system, denitrification measured as N(2)O production was often less (58 to 100%) than the NO(3) uptake because of incomplete inhibition of N(2)O reduction. The NO(3) formed by nitrification and not immediately denitrified but released to the overlying water, uncoupled nitrification, was calculated both from NO(3) dilution and from changes in NO(3) uptake before and after C(2)H(2) addition. These two approaches gave similar results, with rates ranging between 0 and 8.1 mmol of N m day on the four occasions. Attempts to measure total nitrification activity by the difference between NH(4) fluxes before and after C(2)H(2) addition failed because of non-steady-state NH(4) fluxes. The vertical distribution of denitrification and oxygen consumption was studied by use of N(2)O and O(2) microelectrodes. The N(2)O profiles measured during the experimental steady state were often irregularly shaped, and the buildup of N(2)O after C(2)H(2) was added was much too fast to be described by a simple diffusion model. Only bioturbation by a dense population of infauna could explain these observations. This was corroborated by the relationship between diffusive and total fluxes, which showed that only 19 to 36 and 29 to 62% of the total O(2) uptake and denitrification, respectively, were due to diffusion-reaction processes at the regular sediment surface, excluding animal burrows.  相似文献   

16.
Using a combination of high-performance ion chromatography analysis and kinetic studies, the stereospecificity of myo-inositol hexakisphosphate dephosphorylation by the phytate-degrading enzyme P2 of Escherichia coli was established. High-performance ion chromatography revealed that the phytate-degrading enzyme P2 of E. coli degrades myo-inositol hexakisphosphate by stepwise dephosphorylation via D/L-Ins(1,2,3,4,5)P(5), D/L-Ins(2,3,4,5)P(4), D/L-Ins(2,4,5)P(3) or D/L-Ins(1,2,4)P(3), D/L-Ins(1,2)P(2) or Ins(2, 5)P(2) or D/L-Ins(4,5)P(2) to finally Ins(2)P or Ins(5)P. Kinetic parameters for myo-inositol pentakisphosphate hydrolysis by E. coli and wheat phytase, respectively, showed that the myo-inositol pentakisphosphate intermediate produced either by the phytate-degrading enzyme of wheat or E. coli are not identical. The absolute configuration of the myo-inositol pentakisphosphate isomer produced by the E. coli enzyme was determined by taking into consideration that wheat phytase produces predominantly the D-Ins(1, 2,3,5,6)P(5) isomer (Lim, P.E., Tate, M.E., 1973. The phytases: II. Properties of phytase fraction F(1) and F(2) from wheat bran and the myo-inositol phosphates produced by fraction F(2). Biochim. Biophys. Acta 302, 326-328). The data demonstrate that the phytate-degrading enzyme P2 of E. coli dephosphorylates myo-inositol hexakisphosphate in a stereospecific way by sequential removal of phosphate groups via D-Ins(1,2,3,4,5)P(5), D-Ins(2,3,4,5)P(4), D-Ins(2,4,5)P(3), Ins(2,5)P(2) to finally Ins(2)P (notation 6/1/3/4/5).  相似文献   

17.
The present studies were undertaken to investigate the effect of C-atrial natriuretic peptide (ANP)(4-23) and several peptide fragments containing 12 amino acids from different regions of the cytoplasmic domain of natriuretic peptide receptor (NPR)-C on cell proliferation in the absence or presence of angiotensin (ANG) II, endothelin (ET)-1, and arginine vasopressin (AVP) in A-10 vascular smooth muscle cells (VSMC). The peptide fragments used have either complete G(i) activator sequences K(461)-H(472) (peptide 1) and H(481)-H(492) (peptide 3) or partial G(i) activator sequences R(469)-K(480) (peptide 2) and I(465)-H(472) (peptide Y) with truncated COOH or NH(2) terminus, respectively. The other peptide used had no structural specificity (Q(473)-K(480), peptide X) or was the scrambled peptide control for peptide 1 (peptide Z). ANG II, ET-1 and AVP significantly stimulated DNA synthesis in these cells as determined by [(3)H]thymidine incorporation that was inhibited by peptides 1, 2, and 3 and not by peptides X, Y, and Z in a concentration-dependent manner, with an apparent K(i) between 1 and 10 nM. In addition, C-ANP(4-23), which interacts with NPR-C, also inhibited DNA synthesis stimulated by vasoactive peptides; however, the inhibition elicited by C-ANP(4-23) was not additive with the inhibition elicited by peptide 1. On the other hand, basal DNA synthesis in these cells was not inhibited by C-ANP(4-23) or the peptide fragments. Furthermore, vasoactive peptide-induced stimulation of DNA synthesis was inhibited by PD-98059 and wortmannin, and this inhibition was potentiated by peptide 1. In addition, peptide 1 also inhibited vasoactive peptide-induced phosphorylation of ERK1/2 and AKT and enhanced expression of G(i)alpha proteins. These data suggest that C-ANP(4-23) and small peptide fragments containing 12 amino acids irrespective of the region of the cytoplasmic domain of NPR-C inhibit proliferative responses of vasoactive peptides through G(i)alpha protein and MAP kinase/phosphatidylinositol 3-kinase/AKT pathways.  相似文献   

18.
Chromate resistant and reducing strains were isolated from chromium contaminated soil and identified as Bacillus sp. (KCH2 and KCH3), Leucobacter sp. (KCH4) and Exiguobacterium sp. (KCH5). KCH3 and KCH4 showed higher Cr(VI) tolerance (2 mM) and Cr(VI) reduction (1.5 mM) than KCH5 (1.5 mM and 0.75 mM, respectively). Cr(VI) reduction by CFEs of KCH3 and KCH4 showed NAD(P)H dependence, optimum activity at pH 5.5, low K(m) (45-55 microM) and substrate inhibition by Cr(VI) (>75 microM), whereas that of KCH5 showed NADH dependence, pH optimum at 6.0, high K(m) (200 microM) and no inhibition by Cr(VI). Cr(VI) reduction was optimum at 35 degrees C for CFEs of KCH3 and KCH5 and 30 degrees C for that of KCH3. Cr(VI) reduction by CFEs of all the strains were inhibited by Hg(2+) and enhanced by Cu(2+). Activity enhancement by Cu(2+) was more predominant (290%) for KCH4. The characterization of Cr(VI) reduction by CFEs of chromate resistant isolates of different genera is useful for development of Cr(VI) bioremediation.  相似文献   

19.
We showed, using the method of lysis of fibrin plates and five substrate proteins in a thin layer of agar gel, that inorganic orthophosphate (0.001-0.06 M) enhances by 50-250% the activatory functions of streptokinase, urokinase, and tissue plasminogen activator and, in general, by 1.2-12.0 times enhances protein lysis by trypsin, alpha-chymotrypsin, subtilisin, papain, bacterial metalloprotease, and even pepsin at a concentration < 4 mM. At higher concentrations, phosphate sharply inhibited pepsin activity and inhibited by 40-50% gelatin lysis by papain and gelatin (at a peak concentration) and casein lysis by metalloprotease. Inorganic pyrophosphate ions at concentrations of 10(-8)-10(-1) M enhanced the cleavage of a number of proteins by serine proteases and, at concentrations of 10(-5) -10(-3) M, the activities of pepsin, plasminogen tissue activator, and streptokinase by 100 and 40%, respectively. The pyrophosphate concentrations of > 10(-3) and >10(-4) M inhibited pepsin- and metalloprotease-induced lysis of virtually all proteins. ATP increased casein lysis by serine proteases, metalloprotease, and pepsin by 20-60% at concentration of 10(-3) M and by 30-260% at 10(-2) M concentration. At concentrations of 10-2 M, it inhibited the cleavage of some proteins by trypsin, chymotrypsin, papain, and metalloprotease by 20-100%, and, at concentrations of 10(-3) M, lysis of albumin with pepsin and other proteins (except for fibrinogen) by metalloprotease. A GTP concentration of 10(-7)-10(-2) M increased protein degradation by serine proteases, papain, and gelatin lysis by pepsin by 20-90%, whereas albumin lysis was inhibited by 40-70%. The presence of 10(-6)-10(-5) M GTP led to a slightly increased degradation of hemoglobin and casein by bacterial metalloprotease, while 10(-3) M GTP induced a drop in the activity of the metalloprotease by 20-50%. ADP could enhance gelatin lysis by trypsin, casein lysis by pepsin and papain, and inhibited metalloprotease activity by 20-100% (at 10(-3) M). Peculiarities of the effects of AMP and GD(M)P on gelatin lysis were found.  相似文献   

20.
Previous studies proposed that N-ethylmaleimide (NEM) alkylates 3 classes of thiols on skeletal muscle ryanodine receptors (RyRs) producing 3 phases of channel modification, as function of time and concentration. NEM (5 mm) decreased, increased, and then decreased the open probability (P(o)) of the channel by thiol alkylation, a reaction not reversed by reducing agents. We now show that low NEM concentrations (20-200 microm) elicit Ca(2+) release from sarcoplasmic reticulum (SR) vesicles, but contrary to expectations, the effect was fully reversed by reducing agents or by washing SR vesicles. In bilayers, NEM (0.2 mm) increased P(o) of RyRs within seconds when added to the cis (not trans) side, and dithiothreitol (DTT; 1 mm) decreased P(o) in seconds. High (5 mm) NEM concentrations elicited SR Ca(2+) release that was not reversed by DTT, as expected for an alkylation reaction. A non-sulfhydryl reagent structurally related to NEM, N-ethylsuccinimide (0.1-0.5 mm), also elicited SR Ca(2+) release that was not reversed by DTT (1 mm). Other alkylating agents elicited SR Ca(2+) release, which was fully (N-methylmaleimide) or partially (iodoacetic acid) reversed by DTT and inhibited by ruthenium red. Nitric oxide (NO) donors at concentrations that did not activate RyRs inhibited NEM-induced Ca(2+) release, most likely by an interaction of NO with NEM rather than an inactivation of RyRs by NO. Thus, at low concentrations, NEM does not act as a selective thiol reagent and activates RyRs without alkylating critical thiols indicating that the multiple phases of ryanodine binding are unrelated to RyR activity or to NEM alkylation of RyRs.  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