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.
We synthesized vanadyl (oxidation state +IV) and vanadate (oxidation state +V) complexes with the same hydroxamic acid derivative ligand, and assessed their glucose-lowering activities in relation to the vanadium biodistribution behavior in streptozotocin-induced diabetic mice. When the mice received an intraperitoneal injection of the complexes, the vanadate complex more effectively lowered the elevated glucose levels compared with the vanadyl one. The glucose-lowering effect of the vanadate complex was linearly related to its dose within the range from 2.5 to 7.5 mg V/kg. In addition, pretreatment of the vanadate complex induced a larger insulin-enhancing effect than the vanadyl complex. Both complexes were more effective than the corresponding inorganic vanadium compounds. The vanadyl and vanadate complexes, but not the inorganic vanadium compounds, resulted in almost the same organ vanadium distribution. Consequently, the observed differences in the insulin-like activity between the complexes would reflect the potency of the two compounds in the +IV and +V oxidation states in the subcellular region.  相似文献   

2.
The effect of vanadium oxides on living systems may involve the in vivo conversion of vanadate and vanadyl ions. The addition of 5 mM orthovanadate (VO4(3-), V(V)), a known inhibitor of the (Na,K)-ATPase, to yeast cells stopped growth. In contrast, the addition of 5 mM vanadyl (VO2+, V(IV) stimulated growth. Orthovanadate addition to whole cells is known to stimulate various cellular processes. In yeast, both ions inhibited the plasma membrane Mg2+ ATPase and were transported into the cell as demonstrated with [48V]VO4(3-) and VO2+. ESR spectroscopy has been used to measure the cell-associated paramagnetic vandyl ion, while 51V NMR has detected cell-associated diamagnetic vanadium (e.g. V(V)). Cells were exposed to both toxic (5 mM) and nontoxic (1 mM) concentrations of vanadate in the culture medium. ESR showed that under both conditions, vanadate became cell associated and was converted to vanadyl which then accumulated in the cell culture medium. 51V NMR studies showed the accumulation of new cell-associated vanadium resonances identified as dimeric vanadate and decavanadate in cells exposed to toxic amounts of medium vanadate (5 mM). These vanadate compounds did not accumulate in cells exposed to 1 mM vanadate. These studies confirm that the inhibitory form of vanadium usually observed in in vitro experiments is vanadate, in one or more of its hydrated forms. These data also support the hypothesis that the stimulatory form of vanadium usually observed in whole cell experiments is the vanadyl ion or one or more of its liganded derivatives.  相似文献   

3.
The mechanism by which cells reduce cytoplasmic vanadium(V) (vanadate) to vanadium(IV) was investigated using the human red cell as a model system. Vanadate uptake by red cells occurs with a rapid phase involving chemical equilibration across the plasma membrane and a slower phase resulting in a high concentration of bound vanadium(IV). The slow phase was inhibited in glucose-starved cells and restored upon addition of glucose indicating an energy requirement for this process. The time course of vanadium(IV) appearance (monitored by EPR spectroscopy of intact cells) paralleled the slow phase of uptake indicating that this phase involves vanadium reduction. The reduction of intracellular vanadate to vanadium(IV) was nearly quantitative after 23 h. The intracellular reduction is not enzymatic, since a similar time course of vanadium reduction and binding to hemoglobin was observed when glutathione was added to a hemoglobin + vanadate solution in vitro. Vanadium(IV) binding to hemoglobin was reduced by addition of ATP, 2,3-diphosphoglycerate or EDTA, probably through chelation of the cation. The stability constant of the ATP-vanadium (IV) complex was determined to be 150 M-1 at pH 4.9. The time course of red cell vanadate uptake and reduction was followed in the concentration range in which approximately 60% inhibition of the (Na+ + K+)-ATPase is observed. It is concluded that vanadate is reduced by cytoplasmic glutathione in this concentration range and that the reduction explains the resistance of the (Na+ + K+)-ATPase to vanadium in intact cells.  相似文献   

4.
Tunicates, or sea squirts, are known to sequester vanadium to very high concentrations within specialized blood cells. They selectively accumulate the element from seawater against a 106- to 107-fold concentration gradient, and store it mainly as V(III). The mechanism for this selective accumulation involves the facilitated diffusion of vanadate across the blood cell plasma membrane followed by intracellular reduction to a non-transportable cation. Evidence for this mechanism was obtained by studying vanadate and [48V]vanadate influx into living blood cells (vanadocytes). Influx of [48V]vanadate into the cells is a rapid ( ) process which can be saturated (Km = 1.4 (±2%) mM). Net vanadate accumulation is equal to isotopic influx, and accumulated vanadate is not released by washing cells with EDTA. Uncouplers of oxidative phosphorylation and glycolytic inhibitors have no effect on the rate of influx. Phosphate competes with vanadate for transport, and is itself taken up by the cell. The similar anions, sulfate and chromate, neither inhibit transport, nor are they taken up by the vanadocyte. Influx is inhibited by those stilbene disulfonate derivatives known to bind specifically to the external transport site of the anion exchange protein in the human erythrocyte membrane. During the influx of vanadate, the electron paramagnetic resonance (EPR) signal of intracellular vanadyl increases, indicating that transported V(V) is reduced upon entering the cell. The EPR signal of the blood cells at room temperature is characteristic of unbound V(IV), in agreement with reports that reduced vanadate is not bound to a protein or other macromolecule in these cells.  相似文献   

5.
Vanadium compounds are known to stimulate the oxidation of NAD(P)H, but the mechanism remains unclear. This reaction was studied spectrophotometrically and by electron spin resonance spectroscopy (ESR) using vanadium in the reduced state (+4, vanadyl) and the oxidized state (+5, vanadate). In 25 mM sodium phosphate buffer at pH 7.4, vanadyl was slightly more effective in stimulating NADH oxidation than was vanadate. Addition of a superoxide generating system, xanthine/xanthine oxidase, resulted in a marked increase in NADH oxidation by vanadyl, and to a lesser extent, by vanadate. Decreasing the pH with superoxide present increased NADH oxidation for both vanadate and vanadyl. Addition of hydrogen peroxide to the reaction mixture did not change the NADH oxidation by vanadate, regardless of concentration or pH. With vanadyl however, addition of hydrogen peroxide greatly enhanced NADH oxidation which further increased with lower pH. Use of the spin trap DMPO in reaction mixtures containing vanadyl and hydrogen peroxide or a superoxide generating system resulted in the detection by ESR of hydroxyl. In each case, the hydroxyl radical signal intensity increased with vanadium concentration. Catalase was able to inhibit the formation of the DMPO--OH adduct formed by vanadate plus superoxide. These results show that the ability of vanadium to act in a Fenton-type reaction is an important process in the vanadium-stimulated oxidation of NADH.  相似文献   

6.
During incubation with vanadyl, Saccharomyces cerevisiae yeast cells were able to accumulate millimolar concentrations of this divalent cation within an intracellular compartment. The intracellular vanadyl ions were bound to low molecular weight substances. This was indicated by the isotropic nature of the electron paramagnetic resonance (EPR) spectra of the respective samples. Accumulation of intracellular vanadyl was dependent on presence of glucose during incubation. It could be inhibited by various di- and trivalent metal cations. Of these cations lanthanum displayed the strongest inhibitory action. If yeast cells were exposed to more than 50 microM vanadyl sulfate at a pH higher than 4.0, a potassium loss into the medium was detected. The magnitude of this potassium loss suggests a damage of the plasma membrane caused by vanadyl. Upon addition of vanadate to yeast cells surface-bound vanadyl was detectable after several minutes by EPR. This could be the consequence of extracellular reduction of vanadate to vanadyl. The reduction was followed by a slow accumulation of intracellular vanadium, which could be inhibited by lanthanum or phosphate. Therefore, permeation of vanadyl into the cells can be assumed as one mechanism of vanadium accumulation by yeast during incubation with vanadate.  相似文献   

7.
Vanadyl ion (+4 oxidation state) has been shown to be an effective agent for chemoprotection of cancers in animals. For understanding the mechanism, distribution of vanadium was studied. More vanadium was found to accumulate in the nuclei of the liver of rats when it was given as vanadyl sulfate than when it was given as sodium vanadate (+5 oxidation state). The reactivity of vanadyl ion with DNA was investigated by the DNA cleavage technique and the reaction mechanism by ESR spectroscopy. Incubation of double-strand DNA with vanadyl ion and hydrogen peroxide resulted in marked concentration- and pH-dependent DNA cleavage. Studies by the ESR spin-trap method demonstrated that hydroxyl radicals are generated during the reactions of vanadyl ion with hydrogen peroxide. Thus the antineoplastic action of vanadyl ion is proposed to be due to DNA cleavage by hydroxyl radicals generated in the cells.  相似文献   

8.
Vanadium in the 4+ (vanadyl-ion) and 5+ (vanadate-ion) oxidation state stimulates furosemide-sensitive electrogenic Cl- secretion in isolated epithelia of rabbit descending colon. This effect is associated with an increased release of prostaglandin E2 from the tissue. Inhibitors of phospholipase A2 or cyclooxygenase abolish both vanadium-induced release of prostaglandin E2 and Cl- secretion. Neuronal mechanisms are not likely to be involved, as tetrodotoxin does not affect the vanadate induced Cl- secretion. Although vanadate is known to inhibit Na+,K(+)-ATPase activity, no inhibition of active Na+ transport was observed in intact colonic epithelia suggesting a rapid intracellular reduction of vanadate ions to vanadyl ions which have no inhibitory effect on the Na+,K(+)-ATPase. The present findings therefore indicate that vanadate stimulated colonic Cl- secretion involves intracellular conversion of vanadate to vanadyl and release of prostaglandin E2.  相似文献   

9.
K-edge X-ray absorption and EPR spectroscopies were used to test the variation in blood cell vanadium between and within specimens of the tunicate Ascidia ceratodes from Bodega Bay, California. Intracellular vanadium was speciated by fitting the XAS spectra of whole blood cells with linear combinations of the XAS spectra of models. Blood cell samples representing one specimen each, respectively, revealed 92.5 and 38.7% of endogenous vanadium as [V(H(2)O)(6)](3+), indicating dissimilar distributions. Conversely, vanadium distributions within blood cell samples respectively representing one and six specimens proved very similar. The derived array of V(III) complexes was consistent with multiple intracellular regions that differ both in pH and c(sulfate), both within and between specimens. No systematic effect on vanadium distribution was apparent on mixing blood cells. EPR and XAS results indicated at least three forms of endogenous vanadyl ion, two of which may be dimeric. An inverse linear correlation was found between soluble and complexed forms of vanadyl ion, implying co-regulation. The EPR A value of endogenous vanadyl ion [A(0)=(1.062+/-0.008)x10(-2) cm(-1)] was marginally different from that representing Monterey Bay A. ceratodes [A(0)=(1.092+/-0.006) x10(-2) cm(-1)]. Comparisons indicate that Bodega Bay A. ceratodes maintain V(III) in a more acidic intracellular environment on average than do those from Monterey Bay, showing variation across populations. Blood cell vanadium thus noticeably diverges at all organismal levels among A. ceratodes.  相似文献   

10.
Vanadium compounds are shown to have a mitogenic effect on fibroblast cells. The effects of vanadate, vanadyl and pervanadate on the proliferation and morphological changes of Swiss 3T3 cells in culture are compared. Vanadium derivatives induced cell proliferation in a biphasic manner, with a toxic-like effect at doses over 50mM, after 24h of incubation. Vanadyl and vanadate were equally potent at 2.5–10mM. At 50mM vanadate inhibited cell proliferation, whereas slight inhibition was observed at 100mM of vanadyl. At 10mM pervanadate was as potent as vanadate and vanadyl in stimulating fibroblast proliferation, but no effect was observed at lower concentrations. A pronounced cytotoxic-like effect was induced by pervanadate at 50mM. All of these effects were accompanied by morphological changes: transformation of fibroblast shape from polygonal to fusiform; retraction with cytoplasm condensation; and loss of lamellar processes. The magnitude of these transformations correlates with the potency of vanadium derivatives to induce a cytotoxic-like effect: pervanadate>vanadate>vanadyl. These data suggest that the oxidation state and coordination geometry of vanadium determine the degree of the cytotoxicity.  相似文献   

11.
The binuclear iron cluster of uteroferrin in its reduced and enzymatically active pink form is sensitive to a variety or perturbants. Orthophosphate, in the presence or absence of oxygen, rapidly shifts the absorption maximum of pink uteroferrin from 510 to 545 nm, concurrently abolishing the protein's g'av = 1.74 EPR signal. Apparently, therefore, dioxygen is not required for phosphate-induced oxidation of the pink protein's ferrous iron. Pyrophosphate and arsenate produce changes which differ only in degree from those induced by phosphate, suggesting that all of these structurally similar competitive inhibitors bind to a common site. Molybdate, an inhibitor even more potent than phosphate, quantitatively converts the rhombic EPR signal of pink uteroferrin into an axial signal that remains invariant to subsequent additions of phosphate. Thus, there can be inhibition without oxidation, as further evidenced by the complex EPR spectrum of undiminished intensity produced by sulfate. Fluoride, too, induces an axial component in the EPR signal of pink uteroferrin, but at high concentration abolishes the signal entirely. Vanadate also drives the protein to its oxidized, EPR-silent state, serving as an electron acceptor itself to yield the characteristic g' = 2 signal of the vanadyl (VO2+) cation. Remarkably, however, the protein remains pink, demonstrating a dissociation between color and oxidation state. Guanidinium, in contrast, causes a sizeable red shift in the pink protein's absorption maximum without loss of EPR signal intensity, showing dissociation of color and oxidation state in a complementary way.  相似文献   

12.
Vanadium associates with serum transferrin of rats administered vanadyl(IV) sulfate or ammonium metavanadate(V) by gastric intubation. Low molecular weight species account for only 3% of the vanadium present in plasma. The element distributes between the two major isotransferrins in proportion to their concentrations. Rat apotransferrin binds both vanadium(IV) and vanadium(V), forming 2:1 metal-protein complexes in both instances. Although the two isotransferrins apparently differ in their physiological properties, they exhibit identical vanadyl(IV) (VO2+) EPR spectra, indicating identical or very similar metal binding sites for both proteins. In contrast to other transferrins, the two sites of the rat protein are spectroscopically indistinguishable and exhibit a VO2+ EPR spectrum similar to that of the C-terminal metal binding site of human serum transferrin. VO2+ EPR signals are observed with liver, spleen, and kidney tissue samples from animals maintained on a vanadium-supplemented diet. These signals arise from a specific intracellular VO2+ complex with the iron storage protein ferritin.  相似文献   

13.
Microbial reduction and precipitation of vanadium by Shewanella oneidensis   总被引:3,自引:0,他引:3  
Shewanella oneidensis couples anaerobic oxidation of lactate, formate, and pyruvate to the reduction of vanadium pentoxide (V(V)). The bacterium reduces V(V) (vanadate ion) to V(IV) (vanadyl ion) in an anaerobic atmosphere. The resulting vanadyl ion precipitates as a V(IV)-containing solid.  相似文献   

14.
《Inorganica chimica acta》1988,152(3):177-180
ESR spectra due to the vanadyl ion (VO2+, +4 oxidation state) was detected in the branchial basket of Ascidia ahodori, which is reported to contain vanadium in high amounts. The branchial basket, washed with a medium containing 1 mM EDTA, and the supernatant showed different types of vanadyl ESR spectra. On further treatment with 100 mM EDTA the branchial basket gave a characteristic ESR spectrum, indicating that the vanadyl ion binds to a high molecular weight matrix, such as proteins, which makes up the branchial basket. Judging from the relationship of the ESR parameters, g versus A, the vanadyl ion is assumed to ligate with moieties such as deprotonated hydroxyl, or nitrogenous or thiolato groups from oxy- or thiolamino acid residues. The branchial basket was shown to have the ability to reduce added vanadate ion (+5 oxidation state) to the vanadyl form. On the basis of these observations, participation of the branchial basket in vanadium-accumulation by ascidians from seawater is suggested.  相似文献   

15.
Although the number of papers about "vanadium" has doubled in the last decade, the studies about "vanadium and actin" are scarce. In the present review, the effects of vanadyl, vanadate and decavanadate on actin structure and function are compared. Decavanadate (51)V NMR signals, at -516 ppm, broadened and decreased in intensity upon actin titration, whereas no effects were observed for vanadate monomers, at -560 ppm. Decavanadate is the only species inducing actin cysteine oxidation and vanadyl formation, both processes being prevented by the natural ligand of the protein, ATP. Vanadyl titration with monomeric actin (G-actin), analysed by EPR spectroscopy, reveals a 1:1 binding stoichiometry and a K(d) of 7.5 μM(-1). Both decavanadate and vanadyl inhibited G-actin polymerization into actin filaments (F-actin), with a IC(50) of 68 and 300 μM, respectively, as analysed by light scattering assays, whereas no effects were detected for vanadate up to 2 mM. However, only vanadyl (up to 200 μM) induces 100% of G-actin intrinsic fluorescence quenching, whereas decavanadate shows an opposite effect, which suggests the presence of vanadyl high affinity actin binding sites. Decavanadate increases (2.6-fold) the actin hydrophobic surface, evaluated using the ANSA probe, whereas vanadyl decreases it (15%). Both vanadium species increased the ε-ATP exchange rate (k = 6.5 × 10(-3) s(-1) and 4.47 × 10(-3) s(-1) for decavanadate and vanadyl, respectively). Finally, (1)H NMR spectra of G-actin treated with 0.1 mM decavanadate clearly indicate that major alterations occur in protein structure, which are much less visible in the presence of ATP, confirming the preventive effect of the nucleotide on the decavanadate interaction with the protein. Putting it all together, it is suggested that actin, which is involved in many cellular processes, might be a potential target not only for decavanadate but above all for vanadyl. By affecting actin structure and function, vanadium can regulate many cellular processes of great physiological significance.  相似文献   

16.
The direct effect of different vanadium compounds on acid phosphatase (ACP) activity was investigated. Vanadate and vanadyl but not pervanadate inhibited the wheat germ ACP activity. These vanadium derivatives did not alter the fibroblast Swiss 3T3 soluble fraction ACP activity. Using inhibitors of tyrosine phosphatases (PTPases), the wheat germ ACP was partially characterized as a PTPase. This study suggests that the inhibitory ability of different vanadium derivatives to modulate ACP activity seems to depend on the geometry around the vanadium atom more than on the oxidation state. Our results indicate a correlation between the PTPase activity and the sensitivity to vanadate and vanadyl cation.  相似文献   

17.
Biotechnological leaching has been proposed as a suitable method for extraction of vanadium from spent catalysts and oil ash. In the biological leaching process, the vanadium(V) can be reduced to vanadium(IV), which is a less toxic and more soluble form of the vanadium. The present investigation showed that Acidithiobacillus ferrooxidans efficiently reduced vanadium(V) in the form of vanadium pentaoxide, to vanadyl(IV) ions, and tolerated high concentrations of vanadium(IV) and vanadium(V). A. ferrooxidans was compared with Acidithiobacillus thiooxidans, which has previously been utilized for vanadium leaching and reduction. Vanadium pentaoxide and sodium vanadate were used as model compounds. The results of this study indicate possibilities to develop an economical and technically feasible process for biotechnological vanadium recovery.  相似文献   

18.
C R Cremo  G T Long  J C Grammer 《Biochemistry》1990,29(34):7982-7990
The heavy chain of myosin's subfragment 1 (S1) was cleaved at two distinct sites (termed V1 and V2) after irradiation with UV light in the presence of millimolar concentrations of vanadate and in the absence of nucleotides or divalent metals. The V1 site cleavage appeared to be identical with the previously described active site cleavage at serine-180, which is effected by irradiation of a photomodified form of the S1-MgADP-Vi complex [Cremo, C. R., Grammer, J. C., & Yount, R. G. (1989) J. Biol. Chem. 264, 6608-6011]. The V2 site was cleaved specifically, without cleavage at the V1 site, first by formation of the light-stable S1-Co2+ADP-Vi complex at the active site [Grammer, J. C., Cremo, C. R., & Yount, R. G. (1988) Biochemistry 27, 8408-8415] and then by irradiation in the presence of millimolar vanadate. By gel electrophoresis, the V2 site was localized to a region about 20 kDa from the COOH terminus of the S1 heavy chain. From the results of tryptic digestion experiments, the COOH-terminal V2 cleavage peptide appeared to contain lysine-636 in the linker region between the 50- and 20-kDa tryptic peptides of the heavy chain. This site appeared to be the same site cleaved by irradiation of S1 (not complexed with Co2+ADP-Vi) in the presence of millimolar vanadate as previously described [Mocz, G. (1989) Eur. J. Biochem. 179, 373-378]. Cleavage at the V2 site was inhibited by Co2+ but was not significantly affected by the presence of nucleotides or Mg2+ ions. Tris buffer significantly inhibited V2 cleavage. From the results of UV-visible absorption, 51V NMR, and frozen-solution EPR spectral experiments, it was concluded that irradiation with UV light reduced vanadate +5 to the +4 oxidation state, which was then protected from rapid reoxidation by O2 by complexation with the Tris buffer. The relatively stable reduced form or forms of vanadium were not competent to cleave S1 at either the V1 or the V2 site. 51V NMR titration experiments indicated that a tetrameric species of vanadium preferentially bound to S1 and to the S1-MgADP-Vi complex, whereas no binding of either the monomeric or dimeric species could be detected. These results suggest that the vanadate tetramer was responsible for the photocleavage of S1 which occurred at both the V1 and V2 sites in the absence of nucleotides or divalent metals.  相似文献   

19.
We have recently found that both vanadate and vanadyl inhibit ATP-dependent succinyl-CoA synthetase (A-SCS) solubilized from the rat brain mitochondria. Aim of the present study was to estimate a proportion of A-SCS to G-SCS in adult and 5-day-old rat brain and their susceptibility to vanadium ions. The G-SCS to A-SCS ratio of 5-day-old brains was by 196% higher than that in adults. This is in accordance with previous observation that G-SCS is high in tissues metabolizing ketone bodies. Both G-SCS and A-SCS differ in their susceptibility towards vanadium ions. A-SCS of adult brain was more sensitive to vanadate (IC 50 1.6.10(-5) mol.l-1) than was G-SCS (IC 50 6.2.10(-5) mol.l-1). On the contrary G-SCS was more sensitive to vanadyl (IC 50 3.5.10(-4) mol.l-1) than was A-SCS (IC 50 9.0.10(-4) mol.l-1). Also autophosphorylation of G-SCS a-subunit was more resistant to vanadate than A-SCS. In contrast to the adult SCS forms, almost equal susceptibility of A-SCS and G-SCS to vanadyl and vanadate was observed in infant brains. The results suggest some structural (functional) differences between two SCS forms in adults and also between infant and adult G-SCS.  相似文献   

20.
Vanadate ions are shown to inhibit horseradish, squash, and rat intestinal peroxidases by following the reaction spectrophotometrically in a wide range of vanadate concentrations. I50 in phosphate buffer were 43, 9.4, and 535 μM, respectively. No inhibitory effect was found on cow milk lactoperoxidase and beef liver catalase. Gel filtration of peroxidases in the presence of vanadate, as carried out by radioactive48V for horseradish peroxidases (either in aerobic or anoxic conditions) and neutron activation analysis (NAA) for squash peroxidase, demonstrated a binding of vanadium to these enzymes in stoichiometric amounts. Electron paramagnetic resonance spectra of the eluted peaks for the former peroxidase indicated that vanadium is in the +5 oxidation state, but an equilibrium between V (V) and V (IV) in the assay conditions cannot be discarded. Although the inhibitory mechanism remains obscure, some hypotheses are considered. The potential implications that the inhibitory effect of vanadium might have on plant and animal metabolism are also 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