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.
Fits to the vanadium K-edge X-ray absorption spectra (XAS) of five whole blood cell samples from the tunicate Phallusia nigra revealed unprecedented forms of intracellular vanadium. Endogenous vanadium was divided between the V(III) ion (74.2+/-5.1% of total V) and the vanadyl ion [V(IV)=O](2+) (25.2+/-5.4% of total V). The V(III) fraction included both [V(H(2)O)(6)](3+) (36.7+/-5.5%) modeled as VCl(3) in 1 M HCl, and three previously unprecedented chelated V(III) forms (37.5+/-4.6%). Two of these could be represented by the model ligand environments V(acetylacetonate)(3) (17.9+/-3.2%) and K(3)V(catecholate)(3) (13.1+/-4.7%), implying DOPA-like complexation. The third chelated form was represented by the 7-coordinate N(2)O(5) complex Na[V(edta)(H(2)O)] (8.0+/-1.8%). This coordination array, suggestive of a novel mononuclear V(III) protein site, contributed only to fits to samples 1, 2, 3 and 5, which were prepared in the presence of DTT. Endogenous V(IV) (25.2+/-5.4%) was principally modeled as VOCl(2) in 1 M HCl. EPR spectra (averages: A(parallel)=(1.842+/-0.006)x10(-2) cm(-1); A( perpendicular)=(0.718+/-0.007)x10(-2) cm(-1); g(parallel)=1.936+/-0.002; g( perpendicular)=1.990+/-0.001) confirmed the predominance of the aquated vanadyl ion. Blood cell sample five uniquely required the XAS spectrum of VOSO(4) in 0.1 M H(2)SO(4) solution (13.0%) and of [OV(V)(pivalate)(3)] (3.1%) to successfully fit the XAS pre-edge energy region. This endogenous V(V) signal is also unprecedented. These results are compared with those of analogous fits to the blood cells of Ascidia ceratodes and may support assignment of P. nigra to a different genus.  相似文献   

2.
Vanadium K-edge X-ray absorption spectroscopy (XAS) has been used to track the uptake and fate of VO(2+) ion in blood cells from Ascidia ceratodes, following exposure to dithiothreitol (DTT) or to DTT plus VO(2+). The full range of endogenous vanadium was queried by fitting the XAS of blood cells with the XAS spectra of model vanadium complexes. In cells exposed only to DTT, approximately 0.4% of a new V(III) species was found in a site similar to Na[V(edta)(H(2)O)]. With exposure to DTT and VO(2+), average intracellular [VO(aq)](2+) increased from 3% to 5%, and 6% of a new complexed form of vanadyl ion appeared evidencing a ligand array similar to [VO(edta)](2-). At the same time, the relative ratio of blood cell [V(H(2)O)(6)](3+) increased at the expense of [V(H(2)O)(5)(SO(4))](+) in a manner consistent with a significant increase in endogenous acidity. In new UV/Visible experiments, VO(2+) could be reduced to 7-coordinate [V(nta)(H(2)O)(3)] or [V(nta)(ida)](2-) with cysteine methyl ester in pH 6.5 solution. Ascorbate reduced [VO(edta)](2-) to 7-coordinate [V(edta)(H(2)O)](-), while [VO(trdta)](2-) was unreactive. These results corroborate the finding that the reductive EMF of VO(2+) is increased by the availability of a 7-coordinate V(III) product. Finally, a new and complete hypothesis is proposed for an ascidian vanadate reductase. The structure of the enzyme active site, the vanadate-vanadyl-vanadic reduction mechanism, the cellular locale, and elements of the regulatory machinery governing the biological reduction of vanadate and vanadyl ion by ascidians are all predicted. Together these constitute the new field of vanadium redox enzymology.  相似文献   

3.
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   

4.
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   

5.
A series of 2-alkyl-3-hydroxy-4-pyrone oxovanadium(IV) compounds has been synthesized, characterized, and tested for bioactivity as potential insulin-enhancing agents. The vanadyl complexes, bis(maltolato)oxovanadium(IV), BMOV, bis(ethylmaltolato)oxovanadium(IV), BEOV, and bis(isopropylmaltolato)oxovanadium(IV), BIOV, were compared against vanadyl sulfate for glucose-lowering ability, when administered i.p. to STZ-diabetic rats, at a one-time dose of 0.1 mmol kg(-1)body weight. Blood levels of vanadium were determined at regular intervals, to 72 h, following i.p. injection. All complexes tested exceeded vanadyl sulfate in glucose-lowering ability; this effect was not correlated, however, with blood vanadium levels. Analysis of the pharmacokinetics of the disappearance of [ethyl-1-(14)C]BEOV after an oral gavage dose (50 mg kg(-1), 0.144 mmol kg(-1), in a 10 mL kg(-1) volume of 1% CMC solution) indicated clearly that metal ion-ligand dissociation took place relatively soon after oral ingestion of the complex. Half-lives of fast phase uptake and slow phase disappearance for (14)C and V were calculated from a two-compartment model for whole blood, plasma, liver, kidney, bone, small intestine, and lung, ranging from 17 min ( t(1/2)alpha for (14)C, liver) to 30 days ( t(1/2)beta for V, bone). Curves of disappearance of plasma and whole blood (14)C and V diverged dramatically within the first hour after administration of the vanadium complex.  相似文献   

6.
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   

7.
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   

8.
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   

9.
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   

10.
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   

11.
Vanadium binding substance has been partially purified through chromatographies on Sephadex G-25 and SE-Cellulose at pH 2.3. The binding substance was colorless, relatively stable and maintained vanadium ion. The vanadium ion in the substance existed in vanadyl form (VO(IV)). Furthermore, the substance had an apparent affinity for exogenous vanadium ion(V) and contained a reducing sugar.  相似文献   

12.
The effect of the vanadium complex bis[curcumino]oxovanadium (BCOV) on blood glucose level, serum lipid levels, blood pressure and vascular reactivity were studied in non-diabetic and streptozotocin-induced diabetic (STZ-diabetic) rats and compared to that of vanadyl sulfate. Blood glucose level, serum lipid levels, and blood pressure were significantly increased in STZ-diabetic rats. Vascular reactivity to various agonists such as noradrenaline and acetylcholine were significantly increased in STZ-diabetic rats. Blood glucose and serum lipid levels were restored to normal in STZ-diabetic animals treated with vanadyl sulfate at a concentration of 0.5 mmol/kg/day (p.o.). However, vanadyl sulfate at a concentration of 0.2 mmol/kg/day (p.o.) did not produce any significant change in blood glucose and lipid levels. There was no significant effect of vanadyl sulfate (0.2 or 0.5 mmol/kg/day) treatment on blood pressure and vascular reactivity in STZ-diabetic rats. Vanadyl sulfate significantly reduced the body weight of non-diabetic and STZ-diabetic rats. Moreover, it also caused severe diarrhea in both groups of animals. Treatment with BCOV (0.05, 0.1 and 0.2mmol/kg/day, p.o.) significantly decreased blood glucose level and serum lipids in STZ-diabetic rats. Furthermore, administration of BCOV to STZ-diabetic rats restored the blood pressure and vascular reactivity to agonists to normal. There was no significant change in the body weight of BCOV treated non-diabetic and STZ-diabetic rats. Diarrhea was not observed in both BCOV treated groups. In conclusion, the present study shows that the vanadium complex BCOV has antidiabetic and hypolipedimic effects. In addition, it improves the cardiovascular complications associated with diabetes.  相似文献   

13.
EPR spectra due to vanadyl species of blood cells of Ascidia ahodori collected from different places in Japan (Ushimado and Asamushi) were measured. The EPR spectral pattern as well as their EPR parameters for these two species were found to be different from each other. EPR analysis correlating g parallel and A parallel indicated that vanadyl ion in the animals from Ushimado was in a much stronger ligand field than those from Asamushi. Vanadyl species in the Ushimado animals was comparable to the stable complexes such as the vanadyl-ATP complex prepared at pH 12.  相似文献   

14.
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   

15.
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   

16.
The insulinomimetic effect of vanadium is the most remarkable and important among its several biological actions. Vanadyl ion (+4 oxidation state of vanadium) and its complexes have been found to normalize the blood glucose levels of both type 1 and 2 diabetic animals. We have developed insulinomimetic vanadyl complexes having different coordination modes, emphasizing the possible usefulness of vanadyl-picolinate [VO(pa)(2)] and its related complexes with the VO(N(2)O(2)) coordination mode. In order to apply these complexes clinically in the future, the relationship between the chemical structure, insulinomimetic action, organ distribution of vanadium, and blood disposition of vanadyl species must be closely investigated. In the present investigation, we studied the blood disposition of the vanadyl-picolinate complexes in healthy rats, and tried to understand comprehensively the relationship between the structures, insulinomimetic activity, and metallokinetic parameters of the complexes, which had been recently prepared and specifically synthesized for the present study, by using an in vivo blood circulation monitoring -- electron spin resonance (BCM-ESR) method for analyzing ESR signals due to paramagnetic metal ions and complexes in the blood in real time. Metallokinetic parameters were estimated based on the blood clearance curves in terms of a two-compartment pharmacokinetic model, and vanadyl species were indicated to be distributed in peripheral tissues and gradually eliminated from the circulating blood, depending on their chemical structures. Vanadyl concentrations in the blood of rats given bis(5-iodopicolinato)oxovanadium(IV) [VO(5ipa)(2)] and bis(3-methylpicolinato)oxovanadium(IV) [VO(3mpa)(2)] with electron-withdrawing and donating groups, respectively, remained significantly higher and longer, due to their slower clearance rates from the blood, than in rats given other complexes, suggesting that the high exposure and long residence of vanadyl species bring about the high normoglyceric effect in diabetic animals. We then examined the relationship between insulinomimetic activity and metallokinetic parameters in the family of VO(pa)(2) for further development of insulinomimetic vanadyl complexes. IC(50), the 50% inhibitory concentration of the complexes on the free fatty acid release from isolated rat adipocytes treated with epinephrine, was found to be sufficiently correlated with metallokinetic parameters such as area under the concentration curve, mean residence time, total clearance, and distribution volume at steady-state. Furthermore, the in vivo antidiabetic activity of the complexes was enhanced with increasing exposure and residence of vanadyl species in the blood of animals. On the basis of these results, we concluded that in vitro insulinomimetic activity, metallokinetic character, and in vivo antidiabetic action of vanadyl-picolinate complexes are closely related to their chemical structures.  相似文献   

17.
The reproductive composition and genetic diversity of populations of the red seaweed Lithothrix aspergillum Gray (O. Corallinales) were studied at three southern California sites (Shaw's Cove and Treasure Island, Laguna Beach; Indian Rock, Santa Catalina Island) and at a fourth site (Bodega Bay) located in northern California. Sexually reproducing populations were confined to southern California. Diploid individuals were numerically dominant over haploid (gametophytic) individuals at all sites. Intertidal and subtidal subpopulations from Shaw's Cove differed in their reproductive profiles. Most intertidal specimens found on emersed surfaces were densely branched, turf-forming, and bore tetrasporangial (68.6%), carposporangial (11.4%), or spermatangial (5.7%) conceptacles, reflecting a sexual life history; none produced asexual bispores. In contrast, 74.3% of the larger, loosely branched subtidal specimens bore bisporangial conceptacles indicative of asexual reproduction. Nearly 70% of the Indian Rock thalli showed no evidence of conceptacle formation. Only asexual, diploid bispore-producing thalli were obtained from the Bodega Bay site. Genetic diversity (mean number of alleles per locus, percent of polymorphic loci, and average expected heterozygosity) of diploid L. aspergillum populations varied with life-history characteristics and geographic location. A total of 30 alleles was inferred from zymograms of 16 loci examined by starch-gel electrophoresis; of these loci, 11 were polymorphic. The genetic diversity of sexual, diploid populations of L. aspergillum (alleles per locus [A/L] = 1.4-1.5; percent polymorphic loci [%P] = 37.5-50.0) was relatively high compared with other red seaweeds. Lowest diversity (A/L = 1.0; %P = 0.0) occurred in the exclusively asexual Bodega Bay population which consisted of genetic clones. All sexual L. aspergillum populations deviated significantly from Hardy-Wein-berg expectations due to lower than expected heterozygosity. Genetic differentiation (Wright's Fstatistic [FST]; Nei's Genetic Distance [D]) among sexually reproducing southern California populations was low (FST= 0.030) on a local scale (ca. 5 km), suggesting high levels of gene flow, but high genetic differention (FST= 0.390 and 0.406) occurred among southern California populations separated by ca. 70 km. Very high genetic differentiation (FST= 0.583–0.683) was obtained between northern and southern California populations separated by 700–760 km. Our genetic and reproductive data suggest that the L. aspergillum population from Bodega Bay is sustained by perennation, vegetative propagation, or asexual reproduction by bispores and may represent an isolated remnant or a population established by a founder event.  相似文献   

18.
The oxidation of D-lactobionic acid by Cr(VI) yields the 2-ketoaldobionic acid and Cr(3+) as final products when a 20-times or higher excess of the aldobionic acid over Cr(VI) is used. The redox reaction takes place through a complex multistep mechanism, which involves the formation of intermediate Cr(IV) and Cr(V) species. Cr(IV) reacts with lactobionic acid much faster than Cr(V) and Cr(VI) do, and cannot be directly detected. However, the formation of CrO(2)(2+), observed by the first time for an acid saccharide/Cr(VI) system, provides indirect evidence for the intermediacy of Cr(IV) in the reaction path. Cr(VI) and the intermediate Cr(V) react with lactobionic acid at comparable rates, being the complete rate laws for the Cr(VI) and Cr(V) consumption expressed by: -d[Cr(VI)]/dt=[k(I)+k(II)[H(+)]][lactobionicacid][Cr(VI)], where k(I)=(4.1+/-0.1) x 10(-3) M(-1) s(-1) and k(II)=(2.1+/-0.1) x 10(-2) M(-2) s(-1); and -d[Cr(V)]/dt=[k(III)[H(+)]+(k(IV)+k(V)[H(+)])[lactobionicacid]] [Cr(V)], where k(III)=(1.8+/-0.1) x 10(-3) M(-1) s(-1), k(IV)=(1.1+/-0.1) x 10(-2) M(-1) s(-1) and k(V)=(1.0+/-0.1) x 10(-2) M(-2) s(-1), at 33 degrees C. The Electron Paramagnetic Resonance (EPR) spectra show that five-co-ordinate oxo-Cr(V) bischelates are formed at pH 1-5 with the aldobionic acid bound to Cr(V) through the alpha-hydroxyacid group.  相似文献   

19.
A new vanadyl complex, bis(5-iodopicolinato)oxovanadium(IV), VO(IPA)2, with a VO(N2O2) coordination mode, was prepared by mixing 5-iodopicolinic acid and VOSO4 at pH 5, with the structure characterized by electronic absorption, IR, and EPR spectra. Introduction of the halogen atom on to the ligand enhanced the in vitro insulinomimetic activity (IC50 = 0.45 mM) compared with that of bis(picolinato)oxovanadium(IV) (IC50 = 0.59 mM). The hyperglycemia of streptozotocin-induced insulin-dependent diabetic rats was normalized when VO(IPA)2 was given by daily intraperitoneal injection. The normoglycemic effect continued for more than 14 days after the end of treatment. To understand the insulinomimetic action of VO(IPA)2, the organ distribution of vanadium and the blood disposition of vanadyl species were investigated. In diabetic rats treated with VO(IPA)2, vanadium was distributed in almost all tissues examined, especially in bone, indicating that the action of vanadium is not peripheral. Vanadyl concentrations in the blood of normal rats given VO(IPA)2 remain significantly higher and longer than those given other complexes because of its slower clearance rate. VO(IPA)2 binds with the membrane of erythrocytes, probably owing to its high hydrophobicity in addition to its binding with serum albumin. The longer residence of vanadyl species shows the higher normoglyceric effects of VO(IPA)2 among three complexes with the VO(N2O2) coordination mode. On the basis of these results, VO(IPA)2 is indicated to be a preferred agent to treat insulin-dependent diabetes mellitus in experimental animals.  相似文献   

20.
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   

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