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.
Reduction of vanadate to vanadyl by a strain of Saccharomyces cerevisiae   总被引:1,自引:0,他引:1  
Three strains of Saccharomyces cerevisiae, SC-1, DBVPG 6173 and DBVPG 6037, were studied for vanadate resistance in complex Sabouraud medium since they did not thrive in different minimal media (yeast nitrogen base with and without amino acids). The strain SC-1 was resistant up to 16 mm of vanadate, whereas the strains DBVPG 6173 and DBVPG 6037 were inhibited by 8 mm and 4 mm vanadate, respectively. The vanadate resistance in strain SC-1 was constitutive and due to the reduction of this oxyanion to vanadyl, which was detected by EPR spectroscopy and visible spectroscopy. The transformation of vanadate to vanadyl took place during the exponential growth phase; 10 mm of vanadate was reduced to vanadyl outside the cells since the oxyanion was not detected in the cell biomass and only a negligible concentration of vanadyl (25 nmoles mg cells dry weight) was found in the biomass. The other two vanadate-sensitive yeast strains only accumulated vanadate and did not reduce the oxyanion to vanadyl.  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.
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   

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.
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   

6.
The fate of vanadate (+5 oxidation state of vanadium) taken up by the red cell was studied using EPR spectroscopy. The appearance of an EPR signal indicated that most of the cytoplasmic vanadate is reduced to the +4 oxidation state with axial symmetry characteristic of vanadyl ions. The signal at 23 degrees C was characteristic of an immobilized system indicating that the vanadyl ions in the cytoplasm are associated with a large molecule. [48V]Vanadium eluted with hemoglobin when the lysate from Na3[48V[O4-treated red cells was passed through a Sephadex G-100 column and rabbit anti-human hemoglobin serum caused a hemoglobin-specific precipitation of 48V when added to the red cell lysate. Both results indicate that hemoglobin is the protein which binds cytoplasmic vanadyl ions. However, neither sodium vanadate nor vanadyl sulfate bound to purified hemoglobin in vitro. Finally, transient kinetics of vanadyl sulfate interaction with the sodium-and potassium-stimulated adenosine triphosphatase showed that the +4 oxidation state of vanadium is less effective than the +5 oxidation state in inhibiting this enzyme. These results indicate that oxidation-reduction reactions in the cytoplasm are capable of relieving vanadate inhibition of cation transport.  相似文献   

7.
 The action of vanadate on intact human erythrocytes was studied by 1H spin echo and 51V NMR spectroscopy as a model for the behaviour of vanadium(V) complexes in experimental diabetes. Vanadate is reduced by the intact erythrocyte at the expense of intracellular glutathione which rapidly depletes from the intracellular volume. Using the blocking agent 4,4′-diisothio-cyanatostilbene-2,2′-disulfonic acid (DIDS), which specifically blocks the anion transporter, vanadate reduction could be inhibited and glutathione depletion arrested. Thus, for the reaction with the intact cell to occur, vanadium(V) must cross the cell wall, possibly via the anion transporter. Nitrofurantoin was used to inhibit glutathione reductase in the erythrocyte suspensions. Under these conditions, treatment of the cells with vanadate induced glutathione oxidation prior to depletion. A study of the reaction of vanadate with haemolysate indicates that, without the influence of the membrane, rapid oxidation of glutathione to glutathione disulfide by the vanadyl cation occurs with no glutathione depletion, and that under these conditions vanadate reduction is incomplete. This study generates a model for the behaviour of vanadium complexes in vivo, providing a basis for the rational design and synthesis of new vanadium-based agents as insulin mimics. In essence, vanadium is transported across the membrane as vanadate(V), is reduced in situ by glutathione, and becomes complexed to a wide range of intracellular binding sites. Exchange reactions between glutathione and sulfhydryl groups present on haemoglobin and membrane lead to the depletion of glutathione from the cytosol. Received: 12 June 1996 / Accepted: 20 January 1997  相似文献   

8.
Vanadate (sodium orthovanadate), an inhibitor of phosphotyrosine phosphatases (PTPs), mimics many of the metabolic actions of insulin in vitro and in vivo. The potential of vanadate to stimulate glucose transport independent of the early steps in insulin signaling prompted us to test its effectiveness in an in vitro model of insulin resistance. In primary rat adipocytes cultured for 18 h in the presence of high glucose (15 mm) and insulin (10(-7) m), sensitivity to insulin-stimulated glucose transport was decreased. In contrast, there was a paradoxical enhanced sensitivity to vanadate of the insulin-resistant cells (EC(50) for control, 325 +/- 7.5 microm; EC(50) for insulin-resistant, 171 +/- 32 microm; p < 0.002). Enhanced sensitivity was also present for vanadate stimulation of insulin receptor kinase activity and autophosphorylation and Akt/protein kinase B Ser-473 phosphorylation consistent with more effective PTP inhibition in the resistant cells. Investigation of this phenomenon revealed that 1) depletion of GSH with buthionine sulfoximine reproduced the enhanced sensitivity to vanadate while preincubation of resistant cells with N-acetylcysteine (NAC) prevented it, 2) intracellular GSH was decreased in resistant cells and normalized by NAC, 3) exposure to high glucose and insulin induced an increase in reactive oxygen species, which was prevented by NAC, 4) EPR (electron paramagnetic resonance) spectroscopy showed a decreased amount of vanadyl (+4) in resistant and buthionine sulfoximine-treated cells, which correlated with decreased GSH and increased vanadate sensitivity, while total vanadium uptake was not altered, and 5) inhibition of recombinant PTP1B in vitro was more sensitive to vanadate (+5) than vanadyl (+4). In conclusion, the paradoxical increased sensitivity to vanadate in hyperglycemia-induced insulin resistant adipocytes is due to oxidative stress and decreased reduction of vanadate (+5) to vanadyl (+4). Thus, sensitivity of PTP inhibition and glucose transport to vanadate is regulated by cellular redox state.  相似文献   

9.
The influence of vanadate on the adsorption properties of Merocyanine 540 (MC540) to UMR cells was studied by means of specrofluorometry. An increment in the fluorescence was observed in the osteoblasts incubated with 0.1 mM vanadate. This effect could be interpreted in terms of vanadate inhibitory effects on aminotraslocase activity. However, vanadate promotes a similar behavior to that found in UMR 106 cells when it was added to lipid vesicles composed of phosphatidylcholine. The effect of vanadium in different oxidation states, such as vanadate(V) and vanadyl(IV) on lipid membrane properties was examined in large unilamellar vesicles by means of spectrofluorometry employing different probes. Merocyanine 540 and 1,6-diphenylhexatriene were used in order to sense the changes at interfacial and hydrophobic core of membranes, respectively. In contrast to vanadate, vanadyl decreased the fluorescence of MC540. Both vanadium compounds slightly perturbed the hydrocarbon core. The results can be interpreted by the specific adsorption of both compounds on the polar head groups of phospholipid and suggest a possible influence of vanadium compounds on the lipid organization of cell membranes.  相似文献   

10.
Radiolabeled vanadium as either vanadyl ion or vanadate ion was injected intravenously into adult beagle dogs, and blood samples were collected at various times up to 48 hr post injection. For each sample, the distribution of vanadium between the cells and the plasma was determined, and the plasma was analyzed by electrophoresis to identify specific vanadium-binding proteins. Initially, vanadyl ion left the bloodstream more rapidly than vanadate, but the rates equalized after about 5 hr. A significant fraction of the vanadium in blood was associated with the cellular component following injection of both forms of vanadium. About 77% of the plasma vanadium was eventually bound by the serum iron transport protein transferrin, regardless of the vanadium species initially injected. For both vanadyl and vanadate, about 30 hr were required to reach the maximum degree of transferrin binding.  相似文献   

11.
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   

12.
Inhibition of (Na,K)-ATPase by tetravalent vanadium   总被引:1,自引:0,他引:1  
Vanadyl, the tetravalent state of vanadium and a divalent cation, VO2+, was a relatively powerful inhibitor of highly purified membrane-bound sodium and potassium ion transport adenosine triphosphatase. The sensitivity of the ATPase activity to vanadyl characteristically correlated positively with the specific activity of the enzyme preparation. Inhibition ranged from nearly complete inhibition at less than 5 microM vanadyl for some of the purest fractions (specific activity approximately 45 mumol/min/mg of protein) to no observable inhibition at 300 microM vanadyl in one crude preparation of the enzyme with a specific activity of 10 mumol/min/mg of protein. The level of free vanadyl was reduced by incubation with these membranes, but this reduction was not sufficient to account for the low sensitivity to vanadyl observed in crude preparations. A reduction in specific activity by partial inactivation of a sensitive preparation by treatment with FeCl3 and ascorbate reduced its sensitivity to vanadyl. Anionic ligands of the enzyme, vanadate or ATP, increased the rate of recovery from inhibition after chelation of free vanadyl. At pH 6.1, the inhibition was characteristically fully reversible (t1/2 approximately 10 min), whereas at pH 8.1 it was stable for hours. The degree and stability of enzyme inhibition by vanadyl increased for several hours during incubation of the vanadyl-enzyme mixture, and at pH 6.1 the properties of the inhibitor itself also changed with time. Preincubation of the ion at that pH for 5 h before addition of the enzyme produced a more stable inhibition. The time- and pH-dependent changes in the degree and stability of enzyme inhibition probably relate to the complex chemistry of the vanadyl ion in solution.  相似文献   

13.
The iron storage protein, ferritin, represents a possible source of iron for oxidative reactions in biological systems. It has been shown that superoxide and several xenobiotic free radicals can release iron from ferritin by a reductive mechanism. Tetravalent vanadium (vanadyl) reacts with oxygen to generate superoxide and pentavalent vanadium (vanadate). This led to the hypothesis that vanadyl causes the release of iron from ferritin. Therefore, the ability of vanadyl and vanadate to release iron from ferritin was investigated. Iron release was measured by monitoring the generation of the Fe2+-fcrrozine complex. It was found that vanadyl but not vanadate was able to mobilize ferritin iron in a concentration dependent fashion. Initial rates. and iron release over 30 minutes. were unaffected by the addition of superoxide dismutase. Glutathione or vanadate added in relative excess to the concentration of vanadyl, inhibited iron release up to 45%. Addition of ferritin at the concentration used for measuring iron release prevented vanddyl-induced NADH oxidation. Vanadyl promoted lipid peroxidation in phospholipid liposomes. Addition of ferritin to the system stimulated lipid peroxidation up to 50% above that with vanadyl alone. Fcrritin alone did not promote significant levels of lipid peroxidation.  相似文献   

14.
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   

15.
《Free radical research》2013,47(1):125-129
The iron storage protein, ferritin, represents a possible source of iron for oxidative reactions in biological systems. It has been shown that superoxide and several xenobiotic free radicals can release iron from ferritin by a reductive mechanism. Tetravalent vanadium (vanadyl) reacts with oxygen to generate superoxide and pentavalent vanadium (vanadate). This led to the hypothesis that vanadyl causes the release of iron from ferritin. Therefore, the ability of vanadyl and vanadate to release iron from ferritin was investigated. Iron release was measured by monitoring the generation of the Fe2+-fcrrozine complex. It was found that vanadyl but not vanadate was able to mobilize ferritin iron in a concentration dependent fashion. Initial rates. and iron release over 30 minutes. were unaffected by the addition of superoxide dismutase. Glutathione or vanadate added in relative excess to the concentration of vanadyl, inhibited iron release up to 45%. Addition of ferritin at the concentration used for measuring iron release prevented vanddyl-induced NADH oxidation. Vanadyl promoted lipid peroxidation in phospholipid liposomes. Addition of ferritin to the system stimulated lipid peroxidation up to 50% above that with vanadyl alone. Fcrritin alone did not promote significant levels of lipid peroxidation.  相似文献   

16.
17.
Effect of vanadate and vanadyl ions on the ATP-dependent succinyl-CoA synthetase (A-SCS) solubilized by Lubrol-PX from the rat brain mitochondria was tested. Vanadate added to the assay medium at 10(-5) mol.l-1 and 10(-4) mol.l-1 concentrations inhibited the enzyme activity by about 50% and 94%, respectively. When the enzyme was solubilized from the mitochondria preincubated with 10(-4) mol.l-1 and 10(-3) mol.l-1 vanadate, the residual inhibitions were 55% and 100% respectively. The vanadyl cation also induced inhibition of the A-SCS activity but the effect was less expressed. At 10(-4) mol.l-1 concentration only 20% inhibition was achieved. The A-SCS solubilized from the mitochondrial subfractions (perikaryal, light and heavy synaptosomal) differed neither in the activity of A-SCS nor in the susceptibility toward action of vanadium ions. A strong dependence of the vanadate inhibition on the concentration of succinate was observed. The above effect (50% inhibition) could be demonstrated only at saturating concentration of succinate (50 mmol.l-1). The mechanism of vanadium ions action as well as differences between vanadate and vanadyl ions effects are discussed.  相似文献   

18.
Vanadate and vanadyl have many insulin-mimetic effects on cellular metabolism and also have been shown to alter cellular Ca2+ fluxes. In this report, vanadate and vanadyl, like insulin, are shown to inhibit the plasma membrane (Ca2+ + Mg2+)-ATPase/Ca2+ transport system as well as Ca2+ transport by endoplasmic reticulum from rat adipocytes. Ca2+ transport by the endoplasmic reticulum was inhibited half-maximally (I50) by vanadate and vanadyl at concentrations of 30 and 33 microM, respectively. Inhibition of the plasma membrane Ca2+ transport by vanadate and vanadyl was less sensitive, with I50 values of 144 and 92 microM, respectively. These I50 values for plasma membrane Ca2+ transport were similar when measured under conditions of calmodulin-stimulated and non-calmodulin-stimulated Ca2+ transport. The predominant effect of both ions on the kinetic parameters of Ca2+ transport was a substantial decrease in the Vmax by 43-46% for both transport systems. An increase in intracellular Ca2+ following the inhibition of the (Ca2+ + Mg2+)-ATPase/Ca2+ pump in the plasma membrane and endoplasmic reticulum by these vanadium ions may result, at least in part, in the observed insulin-mimetic alterations in cellular metabolism.  相似文献   

19.
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   

20.
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   

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