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 15 毫秒
1.
Agonist-stimulated divalent cation entry was studied in fura-2-loaded hepatocytes. In the presence of extracellular Mn2+, the Ca2(+)-mobilizing hormone vasopressin produced a severalfold stimulation of the basal rate of fura-2 fluorescence quenching as a result of Mn2+ influx; this effect was blocked by the presence of Ni2+ in the incubation medium. Half-maximum and maximum stimulation of Mn2+ influx was observed with 0.1 and 0.8 nM vasopressin, respectively. Agonist-stimulated Mn2+ influx was also seen with angiotensin II, ATP, phenylephrine, and the combination of AlCl3 and NaF. The stimulation of Mn2+ influx did not occur immediately after addition of Ca2(+)-mobilizing agents, but was characterized by a latency period of 20-30 s. In contrast to vasopressin, glucagon did not stimulate Mn2+ influx into hepatocytes, but produced both a 3-fold enhancement of the rate of vasopressin-stimulated Mn2+ entry and the abolishment of the latency period. The effects of glucagon were mimicked by forskolin and dibutyryl cAMP. Pretreatment of hepatocytes with pertussis toxin or depolarization of the cells altered neither the basal rate of Mn2+ entry nor the ability of vasopressin to stimulate this rate. Emptying of the inositol 1,4,5-trisphosphate-sensitive Ca2+ store by treatment with 2,5-di-(tert-butyl)-1,4-benzohydroquinone (tBuBHQ) did not enhance Mn2+ entry into hepatocytes; however, exposure of the cells to tBuBHQ for 2 min markedly enhanced the ability of vasopressin, alone or in combination with glucagon, to increase the rate of Mn2+ influx. Furthermore, pretreatment with tBuBHQ for 2 min abolished the latency of vasopressin-stimulated Mn2+ influx. It is concluded that Ca2(+)-mobilizing hormones stimulate Ca2+ influx in hepatocytes, possibly through receptor-operated Ca2+ channels. The stimulation of divalent cation entry is transduced by a G protein, and the rate of influx appears to be controlled both by the intracellular level of cAMP and the empty state of an intracellular Ca2+ pool that may be inositol 1,4,5-trisphosphate-insensitive.  相似文献   

2.
1. The inflow of Mn2+ across the plasma membranes of isolated hepatocytes was monitored by measuring the quenching of the fluorescence of intracellular quin2, by atomic absorption spectroscopy and by the uptake of 54Mn2+. The inflow of other divalent metal ions was measured using quin2. 2. Under ionic conditions which resembled those present in the cytoplasmic space, Mn2+, Zn2+, Co2+, Ni2+ and Cd2+ each quenched the fluorescence of a solution of Ca2(+)-quin2. 3. The addition of Mn2+, Zn2+, Co2+, Ni2+ or Cd2+ to cells loaded with quin2 caused a time-dependent decrease in the fluorescence of intracellular quin2. Plots of the rate of decrease in fluorescence as a function of the concentration of Mn2+ reached a plateau at 100 microM-Mn2+. 4. The rate of decrease in fluorescence induced by Mn2+ was stimulated by 20% in the presence of vasopressin. The effect of vasopressin was completely inhibited by 200 microM-verapamil. Adrenaline, angiotensin II and glucagon also stimulated the rate of decrease in the fluorescence of intracellular quin2 induced by Mn2+. 5. The rate of decrease in fluorescence induced by Zn2+, Co2+, Ni2+ or Cd2+ was stimulated by between 20 and 190% in the presence of vasopressin or angiotensin II. 6. The rates of uptake of Mn2+ measured by atomic absorption spectroscopy or by using 54Mn2+ were inhibited by about 20% by 1.3 mM-Ca2+o and stimulated by 30% by vasopressin. 7. Plots of Mn2+ uptake, measured by atomic absorption spectroscopy or with 54Mn2+, as a function of the extracellular concentration of Mn2+ were biphasic over the range 0.05-1.0 mM added Mn2+ and did not reach a plateau at 1.0 mM-Mn2+. 8. It is concluded that (i) hepatocytes possess both a basal and a receptor-activated divalent cation inflow system, each of which has a broad specificity for metal ions, and (ii) the receptor-activated divalent cation inflow system is the receptor-operated Ca2+ channel.  相似文献   

3.
M E Lee  T Nowak 《Biochemistry》1992,31(7):2172-2180
A new, more gentle enzyme purification for yeast enolase was developed. A series of kinetic experiments was performed with yeast enolase where the concentration of Mg(II) is kept constant and at the Km' level; the addition of Mn(II), Zn(II), or Cu(II) gives a hyperbolic decrease in the enzyme activity. The final velocity of these mixed-metal systems is the same as the velocity obtained only with Mn(II), Zn(II), or Cu(II), respectively. The concentration of the second metal that gives half-maximal effect in the presence of Mg(II) is approximately the same as the apparent Km (Km') value measured for that cation alone. Direct binding of Mn(II) to apoenolase in the absence and presence of Mg(II) shows that Mn(II) and Mg(II) compete for the same metal site on enolase. In the presence of D-2-phosphoglycerate (PGA) and Mg(II), only a single cation site per monomer is occupied by Mn(II). Water proton relaxation rate (PRR) studies of enzyme-ligand complexes containing Mn(II) and Mn(II) in the presence of Mg(II) are consistent with Mn(II) binding at site I under both conditions. PRR titrations of ligands such as the substrate PGA or the inhibitors orthophosphate or fluoride to the enolase-Mn(II)-Mg(II) complex are similar to those obtained for the enolase-Mn(II) complex, also indicating that Mn(II) is at site I in the presence of Mg(II). High-resolution 1H and 31P NMR was used to determine the paramagnetic effect of enolase-bound Mn(II) on the relaxation rates of the nuclei of the competitive inhibitor phosphoglycolate. The distances between the bound Mn(II) and the nuclei were calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
We have used electron spin resonance (ESR) techniques to study the binding of divalent cations to diphtheria toxin (DT). Addition of DT to Mn(II) solution at a stoichiometry of 2:1 (DT:Mn) induces a 79% loss in the intensity of the ESR spectrum of Mn(II) suggesting a strong binding of Mn(II) to DT. Inclusion of Ca(II) at a ratio of 1:2:1 (Ca:DT:Mn) in the reaction mixture restores the intensity of the Mn(II) signal to 64%. This indicates that Ca(II) and Mn(II) share same binding site(s) in DT. The results presented in this communication suggest that DT is a Ca(II) binding protein.  相似文献   

5.
Transition row metal ions are both essential and toxic to microorganisms. Zinc in excess has significant toxicity to bacteria, and host release of Zn(II) at mucosal surfaces is an important innate defence mechanism. However, the molecular mechanisms by which Zn(II) affords protection have not been defined. We show that in Streptococcus pneumoniae extracellular Zn(II) inhibits the acquisition of the essential metal Mn(II) by competing for binding to the solute binding protein PsaA. We show that, although Mn(II) is the high-affinity substrate for PsaA, Zn(II) can still bind, albeit with a difference in affinity of nearly two orders of magnitude. Despite the difference in metal ion affinities, high-resolution structures of PsaA in complex with Mn(II) or Zn(II) showed almost no difference. However, Zn(II)-PsaA is significantly more thermally stable than Mn(II)-PsaA, suggesting that Zn(II) binding may be irreversible. In vitro growth analyses show that extracellular Zn(II) is able to inhibit Mn(II) intracellular accumulation with little effect on intracellular Zn(II). The phenotype of S. pneumoniae grown at high Zn(II):Mn(II) ratios, i.e. induced Mn(II) starvation, closely mimicked a ΔpsaA mutant, which is unable to accumulate Mn(II). S. pneumoniae infection in vivo elicits massive elevation of the Zn(II):Mn(II) ratio and, in vitro, these Zn(II):Mn(II) ratios inhibited growth due to Mn(II) starvation, resulting in heightened sensitivity to oxidative stress and polymorphonuclear leucocyte killing. These results demonstrate that microbial susceptibility to Zn(II) toxicity is mediated by extracellular cation competition and that this can be harnessed by the innate immune response.  相似文献   

6.
The three-dimensional structures of two isozymes of manganese peroxidase (MnP) have been predicted from homology modeling using lignin peroxidase as a template. Although highly homologous, MnP differs from LiP by the requirement of Mn(II) as an intermediate in its oxidation of substrates. The Mn(II) site is absent in LiP and unique to the MnP family of peroxidases. The model structures were used to identify the unique Mn(II) binding sites, to determine to what extent they were conserved in the two isozymes, and to provide insight into why this site is absent in LiP. For each isozyme of MnP, three candidate Mn(II) binding sites were identified. Energy optimizations of the three possible Mn(II) enzyme complexes allowed the selection of the most favorable Mn(II) binding site as one with the most anionic oxygen moieties best configured to act as ligands for the Mn(II). At the preferred site, the Mn(II) is coordinated to the carboxyl oxygens of Glu-35, Glu-39, and Asp-179, and a propionate group of the heme. The predicted Mn(II) binding site is conserved in both isozymes. Comparison between the residues at this site in MnP and the corresponding residues in LiP shows that two of the three anionic residues in MnP are replaced by neutral residues in LiP, explaining why LiP does not bind Mn(II). © 1994 Wiley-Liss, Inc.  相似文献   

7.
Lactobacillus plantarum has an unusually high Mn(II) requirement for growth and accumulated over 30 mM intracellular Mn(II). The acquisition of Mn(II) by L. plantarum occurred via a specific active transport system powered by the transmembrane proton gradient. The Mn(II) uptake system has a Km of 0.2 microM and a Vmax of 24 nmol mg-1 of protein min-1. Above a medium Mn(II) concentration of 200 microM, the intracellular Mn(II) level was independent of the medium Mn(II) and unresponsive to oxygen stresses but was reduced by phosphate limitation. At a pH of 5.5, citrate, isocitrate, and cis-aconitate effectively promoted MN(II) uptake, although measurable levels of 1,5-[14C]citrate were not accumulated. When cells were presented with equimolar Mn(II) and Cd(II), Cd(II) was preferentially taken up by the Mn(II) transport system. Both Mn(II) and Cd(II) uptake were greatly increased by Mn(II) starvation. Mn(II) uptake by Mn(II)-starved cells was subject to a negative feedback regulatory mechanism functioning less than 1 min after exposure of the cells to Mn(II) and independent of protein synthesis. When presented with a relatively large amount of exogenous Mn(II), Mn(II)-starved cells exhibited a measurable efflux of their internal Mn(II), but the rate was only a small fraction of the maximal Mn(II) uptake rate.  相似文献   

8.
9.
Cobalt(II) oxidation in aquatic environments has been shown to be linked to Mn(II) oxidation, a process primarily mediated by bacteria. This work examines the oxidation of Co(II) by the spore-forming marine Mn(II)-oxidizing bacterium Bacillus sp. strain SG-1, which enzymatically catalyzes the formation of reactive nanoparticulate Mn(IV) oxides. Preparations of these spores were incubated with radiotracers and various amounts of Co(II) and Mn(II), and the rates of Mn(II) and Co(II) oxidation were measured. Inhibition of Mn(II) oxidation by Co(II) and inhibition of Co(II) oxidation by Mn(II) were both found to be competitive. However, from both radiotracer experiments and X-ray spectroscopic measurements, no Co(II) oxidation occurred in the complete absence of Mn(II), suggesting that the Co(II) oxidation observed in these cultures is indirect and that a previous report of enzymatic Co(II) oxidation may have been due to very low levels of contaminating Mn. Our results indicate that the mechanism by which SG-1 oxidizes Co(II) is through the production of the reactive nanoparticulate Mn oxide.  相似文献   

10.
Cobalt(II) oxidation in aquatic environments has been shown to be linked to Mn(II) oxidation, a process primarily mediated by bacteria. This work examines the oxidation of Co(II) by the spore-forming marine Mn(II)-oxidizing bacterium Bacillus sp. strain SG-1, which enzymatically catalyzes the formation of reactive nanoparticulate Mn(IV) oxides. Preparations of these spores were incubated with radiotracers and various amounts of Co(II) and Mn(II), and the rates of Mn(II) and Co(II) oxidation were measured. Inhibition of Mn(II) oxidation by Co(II) and inhibition of Co(II) oxidation by Mn(II) were both found to be competitive. However, from both radiotracer experiments and X-ray spectroscopic measurements, no Co(II) oxidation occurred in the complete absence of Mn(II), suggesting that the Co(II) oxidation observed in these cultures is indirect and that a previous report of enzymatic Co(II) oxidation may have been due to very low levels of contaminating Mn. Our results indicate that the mechanism by which SG-1 oxidizes Co(II) is through the production of the reactive nanoparticulate Mn oxide.  相似文献   

11.
Bacteria belonging to the Roseobacter clade of the alpha-Proteobacteria occupy a wide range of environmental niches and are numerically abundant in coastal waters. Here we reveal that Roseobacter-like bacteria may play a previously unrecognized role in the oxidation and cycling of manganese (Mn) in coastal waters. A diverse array of Mn(II)-oxidizing Roseobacter-like species were isolated from Elkhorn Slough, a coastal estuary adjacent to Monterey Bay in California. One isolate (designated AzwK-3b), in particular, rapidly oxidizes Mn(II) to insoluble Mn(III, IV) oxides. Interestingly, AzwK-3b is 100% identical (at the 16S rRNA gene level) to a previously described Pfiesteria-associated Roseobacter-like bacterium, which is not able to oxidize Mn(II). The rates of manganese(II) oxidation by live cultures and cell-free filtrates are substantially higher when the preparations are incubated in the presence of light. The rates of oxidation by washed cell extracts, however, are light independent. Thus, AzwK-3b invokes two Mn(II) oxidation mechanisms when it is incubated in the presence of light, in contrast to the predominantly direct enzymatic oxidation in the dark. In the presence of light, production of photochemically active metabolites is coupled with initial direct enzymatic Mn(II) oxidation, resulting in higher Mn(II) oxidation rates. Thus, Roseobacter-like bacteria may not only play a previously unrecognized role in Mn(II) oxidation and cycling in coastal surface waters but also induce a novel photooxidation pathway that provides an alternative means of Mn(II) oxidation in the photic zone.  相似文献   

12.
D T Lodato  G H Reed 《Biochemistry》1987,26(8):2243-2250
The 2 equiv of divalent cation that are required cofactors for pyruvate kinase reside in sites of different affinities for different species of cation [Baek, Y. H., & Nowak, T. (1982) Arch. Biochem. Biophys. 217, 491-497]. The intrinsic selectivity of the protein-based site for Mn(II) and of the nucleotide-based site for Mg(II) has been exploited in electron paramagnetic resonance (EPR) investigations of ligands for Mn(II) at the protein-based site. Oxalate, a structural analogue of the enolate of pyruvate, has been used as a surrogate for the reactive form of pyruvate in complexes with enzyme, Mn(II), Mg(II), and ATP. Addition of Mg(II) to solutions of enzyme, Mn(II), ATP, and oxalate sharpens the EPR signals for the enzyme-bound Mn(II). Superhyperfine coupling between the unpaired electron spin of Mn(II) and the nuclear spin of 17O, specifically incorporated into oxalate, shows that oxalate is bound at the active site as a bidentate chelate with Mn(II). Coordination of the gamma-phosphate of ATP to this same Mn(II) center is revealed by observation of superhyperfine coupling form 17O regiospecifically incorporated into the gamma-phosphate group of ATP. By contrast, 17O in the alpha-phosphate or in the beta-phosphate groups of ATP does not influence the spectrum. Experiments in 17O-enriched water show that there is also a single water ligand bound to the Mn(II). These data indicate that ATP bridges Mn(II) and Mg(II) at the active site. A close spacing of the two divalent cations is also evident from the occurrence of magnetic interactions for complexes in which 2 equiv of Mn(II) are present at the active site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The oxidation states of the Mn atoms in three derivatives of Mn catalase have been characterized using a combination of X-ray absorption near-edge structure (XANES) and EPR spectroscopies. The as-isolated enzyme has an average oxidation state of Mn(III) and contains a Mn(III) form, together with a reduced Mn(II) form and a variable amount (10-25%) of a Mn(III)/Mn(IV) mixed-valence derivative. Treatment with NH2OH rapidly reduces the majority of the enzyme to a Mn(II) derivative with no loss of activity. Inactivation by treatment with NH2OH + H2O2 converts all of the enzyme to a mixed-valence Mn(III)/Mn(IV) form. The inactive, mixed-valence derivative can be completely reactivated by long-term (greater than 1 h) anaerobic incubation with NH2OH, giving a reduced Mn(II)/Mn(II) derivative. These data suggest a catalytic model in which the enzyme cycles between a reduced Mn(II)/Mn(II) state and an oxidized Mn(III)/Mn(III) state.  相似文献   

14.
Rates of oxidation of Mn(II) were measured by an in situ incubation technique in the water column of Toolik Lake, Alaska. Measured rates were lower than those observed in other aquatic systems but were sufficient to oxidize all Mn(II) in the lake within a 3-month period. Measured rates compared favorably with rates estimated from a previous study of the geochemical cycling of Mn in Toolik Lake. The Mn(II) oxidation was largely microbially mediated, as indicated by inhibition of oxidation rates by sodium azide. Azide had been previously demonstrated to be a suitable microbial poison for studying Mn(II) oxidation in seawater. This study demonstrates that azide is also a suitable poison for freshwaters and that it inhibits microbial but not abiotic oxidation of Mn(II). Manganese(II) oxidation rates were similar during cold, under-ice conditions in early spring and during warmer summer conditions. This observation suggests that Mn(II) concentration, rather than temperature or oxygen concentration, is the most important factor regulating Mn(II) oxidation rates in Toolik Lake.  相似文献   

15.
The mechanism of DNA damage by hydrazine in the presence of metal ions was investigated by DNA sequencing technique and ESR-spin trapping method. Hydrazine caused DNA damage in the presence of Mn(III), Mn(II), Cu(II), Co(II), and Fe(III). The order of inducing effect on hydrazine-dependent DNA damage (Mn(III) greater than Mn(II) approximately Cu(II) much greater than Co(II) approximately Fe(III)) was related to that of the accelerating effect on the O2 consumption rate of hydrazine autoxidation. DNA damage by hydrazine plus Mn(II) or Mn(III) was inhibited by hydroxyl radical scavengers and superoxide dismutase, but not by catalase. On the other hand, bathocuproine and catalase completely inhibited DNA damage by hydrazine plus Cu(II), whereas hydroxyl radical scavengers and superoxide dismutase did not. Hydrazine plus Mn(II) or Mn(III) caused cleavage at every nucleotide with a little weaker cleavage at adenine residues, whereas hydrazine plus Cu(II) induced piperidine-labile sites frequently at thymine residues, especially of the GTC sequence. ESR-spin trapping experiments showed that hydroxyl radical is generated during the Mn(III)-catalyzed autoxidation of hydrazine, whereas hydrogen atom adducts of spin trapping reagents are generated during Cu(II)-catalyzed autoxidation. The results suggest that hydrazine plus Mn(II) or Mn(III) generate hydroxyl free radical not via H2O2 and that this hydroxyl free radical causes DNA damage. A possibility that the hydrogen atom releasing compound participates in hydrazine plus Cu(II)-induced DNA damage is discussed.  相似文献   

16.
Heat treatment of Pedomicrobium sp. ACM 3067 enhanced the adsorption of Mn(II) to whole cells but abolished Mn(II)-oxidising activity. In whole cells, optimal Mn(II)-oxidising activity occurred at pH 7 and 25 °C. The apparent K m of the Mn(II)-oxidising system for Mn(II) was 26 μM. These data confirm that Mn(II) oxidation is an enzymic process in Pedomicrobium sp. ACM 3067. Measurement of Mn(II) oxidation during the growth cycle demonstrated that the highest activity occurred during early- to mid-exponential phase and was independent of the presence of Mn in the growth medium. Mn(II)-oxidising activity was localised to the membrane fraction. Transmission electron microscopy showed that this fraction consisted of double-layered membrane vesicles. Positively charged molecules such as poly-l-lysine interfered with the adsorption and oxidation of Mn(II) by whole cells and membranes. Similarly, aminoglycoside antibiotics such as gentamicin sulfate proved to be potent inhibitors of Mn(II) oxidation. Treatment of cells with the copper chelator diethyldithiocarbamate inhibited Mn(II) oxidation. Enzyme activity was restored by the addition of Cu(II) ions, but not by Co(II) nor Zn(II). We conclude that Mn(II) oxidation in Pedomicrobium sp. ACM 3067 is catalysed by a Cu-dependent enzyme. Received: 14 September 1998 / Accepted: 4 January 1999  相似文献   

17.
18.
Potential mechanisms for the lack of Fe(II) accumulation in Mn(IV)‐con‐taining anaerobic sediments were investigated. The addition of Mn(IV) to sediments in which Fe(III) reduction was the terminal electron‐accepting process removed all the pore‐water Fe(II), completely inhibited net Fe(III) reduction, and stimulated Mn(IV) reduction. In a solution buffered at pH 7, Mn(IV) oxidized Fe(II) to amorphic Fe(III) oxide. Mn(IV) naturally present in oxic freshwater sediments also rapidly oxidized Fe(II). A pure culture of a dissimilatory FE(III)‐ and Mn(FV)‐reducing organism isolated from the sediments reduced Fe(III) to Fe(II) in the presence of Mn(IV) when ferrozine was present to trap Fe(II) before Mn(IV) oxidized it. Depth profiles of dissolved iron and manganese reported in previous studies suggest that Fe(II) diffusing up from the zone of Fe(III) reduction is consumed within the Mn(IV)‐reducing zone. These results demonstrate that preferential reduction of Mn(IV) by Fe(III)‐reducing bacteria cannot completely explain the lack of Fe(II) accumulation in anaerobic, Mn(IV)‐containing sedments, and indicate that Mn(IV) oxidation of Fe(II) is the mechanism that ultimately prevents Fe(II) accumulation.  相似文献   

19.
Previous studies have demonstrated that in glia and astrocytes Mn(II) is distributed with ca. 30–40% in the cytoplasm, 60–70% in mitochondria. Ca(II) ions were observed to alter both the flux rates and distribution of Mn(II) ions in primary cultues of chick glia and rat astrocytes. External (influxing) Ca(II) ions had the greatest effect on Mn(II) uptake and efflux, compared to internal (effluxing) or internal-external equilibrated Ca(II) ions. External (influxing) Ca(II) ions inhibited the net rate and extent of Mn(II) uptake but enhanced Mn(II) efflux from mitochondria. These observations differ from Ca(II)–Mn(II) effects previously reported with brain (neuronal) mitochondria. Overall, increased cytoplasmic Ca(II) acts to block Mn(II) uptake and enhance Mn(II) release by mitochondria, which serve to increase the cytoplasmic concentration of free Mn(II). A hypothesis is presented involving external L-glutamate acting through membrane receptors to mobilize cell Ca(II), which in turn causes mitochondrial Mn(II) to be released. Because the concentration of free cytoplasmic Mn(II) is poised near the Kd for Mn(II) with glutamine synthetase, a slight increase in cytoplasmic Mn(II) will directly enhance the activity of glutamine synthetase, which catalyzes removal of neurotoxic glutamate and ammonia.  相似文献   

20.
An ultrasensitive colorimetric assay for manganese is described. It is based upon the catalysis, by Mn(II), of the photochemical oxidation of o-dianisidine, sensitized by riboflavin. Catalase increases the Mn(II)-catalyzed rate of photosensitized oxidation of dianisidine to the bisazobiphenyl, while superoxide dismutase inhibits the rate. The mechanism appears to involve oxidation of Mn(II) by O2-, followed by oxidation of dianisidine by MnO2+ in equilibrium Mn(III). Cu(II) interferes, but Zn(II), Fe(II), Fe(III), Co(II), and Ni(II) do not. Chelating agents and thiol reductants also interfere. Interference by Cu(II) can be overcome by the addition of cyanide, while interference by organic compounds can be surmounted by wet ashing. This assay provides a linear response to Mn(II) over the range 10-2500 nM. The limit of detection was 5 nM Mn(II).  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