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.
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   

2.
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   

3.
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   

4.
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   

5.
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   

6.
Among vanadium's wide variety of biological functions, its insulin-mimetic effect is the most interesting and important. Recently, the vanadyl ion (+4 oxidation state of vanadium) and its complexes have been shown to normalize the blood glucose levels of streptozotocin-induced diabetic rats (STZ-rats). During our investigations to find more effective and less toxic vanadyl complexes, the vanadyl-methylpicolinate complex (VO-MPA) was found to exhibit higher insulin-mimetic activity and less toxicity than other complexes, as evaluated by both in vitro and in vivo experiments. Electron spin resonance (ESR) is capable of measuring the paramagnetic species in biological samples. We have developed the in vivo blood circulation monitoring-electron spin resonance (BCM-ESR) method to analyze the ESR signals due to stable organic radicals in real time. In the present investigation, we have applied this method to elucidate the relationship between the blood glucose normalizing effect of VO-MPA and the global disposition of paramagnetic vanadyl species. This paper describes the results of vanadyl species in the circulating blood of rats following intravenous administration of vanadyl compounds. ESR spectra due to the presence of vanadyl species were obtained in the circulating blood, and their pharmacokinetic parameters were estimated using compartment models. The results indicate that vanadyl species are distributed considerably to the peripheral tissues, as estimated by BCM-ESR, and eliminated from the body through the urine, as estimated by ESR at 77 K. The exposure of vanadyl species in the blood was found to be enhanced by VO-MPA treatment. Given these results, we concluded that the pharmacokinetic character of vanadyl species is closely related with the structure and antidiabetic activity of the vanadyl compounds.  相似文献   

7.
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   

8.
The binding of vanadyl to a porcine and bovine testicular S-100-like protein and to calmodulin was demonstrated using X-band (9.2 gHz) electron paramagnetic resonance (EPR) spectroscopy in aqueous solution at pH 7.4. In liquid solutions at 22°C, the vanadyl-protein complexes yielded VO2+ near rigid limit spectra. At 122 K, each of the three high-field resonances (i.e., 3/2, 5/2, and 7/2 parallel components) splits into two components indicating the presence of two classes of vanadyl-binding sites in each protein. The spectra of the frozen solutions were simulated to give parallel and perpendicular components of the hyperfine coupling constant and g factors similar to other vanadyl-protein complexes.  相似文献   

9.
The binding of vanadyl to a porcine and bovine testicular S-100-like protein and to calmodulin was demonstrated using X-band (9.2 gHz) electron paramagnetic resonance (EPR) spectroscopy in aqueous solution at pH 7.4. In liquid solutions at 22°C, the vanadyl-protein complexes yielded VO2+ near rigid limit spectra. At 122 K, each of the three high-field resonances (i.e., 3/2, 5/2, and 7/2 parallel components) splits into two components indicating the presence of two classes of vanadyl-binding sites in each protein. The spectra of the frozen solutions were simulated to give parallel and perpendicular components of the hyperfine coupling constant and g factors similar to other vanadyl-protein complexes.  相似文献   

10.
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   

11.
Site-directed mutations D262C, D262H, D262N, and D262T were made to the beta subunit Walker Homology B aspartate of chloroplast F(1)-ATPase in Chlamydomonas. Photoautotrophic growth and photophosphorylation rates were 3-14% of wild type as were ATPase activities of purified chloroplast F(1) indicating that betaD262 is an essential residue for catalysis. The EPR spectrum of vanadyl bound to Site 3 of chloroplast F(1) as VO(2+)-ATP gave rise to two EPR species designated B and C in wild type and mutants. (51)V-hyperfine parameters of species C, present exclusively in the activated enzyme state, did not change significantly by the mutations examined indicating that it is not an equatorial ligand to VO(2+), nor is it hydrogen-bonded to a coordinated water at an equatorial position. Every mutation changed the ratio of EPR species C/B and/or the (51)V-hyperfine parameters of species B, the predominant conformation of VO(2+)-nucleotide bound to Site 3 in the latent (down-regulated) state. The results indicate that the Walker Homology B aspartate coordinates the metal of the predominant metal-nucleotide conformation at Site 3 in the latent state but not in the conformation present exclusively upon activation and elucidates one of the specific changes in metal ligation involved with activation.  相似文献   

12.
As a contribution to the development of novel vanadyl complexes with potential insulin-mimetic activity, three new oxovanadium(IV) complexes with the formula VO(L)(2), where L are 3-amino-quinoxaline-2-carbonitrile N(1),N(4)-dioxide derivatives, have been synthesized. Complexes have been characterized by elemental and thermal analyses, fast atom bombardment mass spectroscopy (FAB-MS), conductivity measurements and electronic, Fourier transform infrared (FTIR) and electron paramagnetic resonance (EPR) spectroscopies. The in vitro insulin-mimetic activity of the vanadyl complexes has been estimated by lipolysis inhibition tests, in which the inhibition of the release of free fatty acid from isolated rat adipocytes treated with epinephrine was determined. All the complexes showed inhibitory effects on free fatty acid release. [V(IV)O(3-amino-6(7)-bromoquinoxaline-2-carbonitrile N(1),N(4)-dioxide)(2)] exhibited higher in vitro insulin-mimetic activity than the very active bis(6-methylpicolinato)oxovanadium(IV), VO(6mpa)(2). This new vanadyl complex is expected to exhibit a higher blood glucose lowering activity than VO(6mpa)(2) in diabetic animals.  相似文献   

13.
Vanadium uptake by whole cells and isolated cell walls of the yeast Saccharomyces cerevisiae was studied. When orthovanadate was added to wild-type S. cerevisiae cells growing in rich medium, growth was inhibited as a function of the VO4 3- concentration and the growth was completely arrested at a concentration of 20 mM of VO4 3- in YEPD. Electron paramagnetic resonance (EPR) spectroscopy was used to obtain structural and dynamic information about the cell-associated paramagnetic vanadyl ion. The presence of EPR signals indicated that vanadate was reduced by whole cells to the vanadyl ion. On the contrary, no EPR signals were detected after interaction of vanadate with isolated cell walls. A mobile and an immobile species associated in cells with small chelates and with macromolecular sites, respectively, were identified. The value of rotational correlation time r indicated the relative motional freedom at the macromolecular site. A strongly immobilized vanadyl species bound to polar sites mainly through coulombic attractions was detected after interaction of VO2+ ions with isolated cell walls.  相似文献   

14.
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   

15.
Nitrous oxide reductase from Wolinella succinogenes, an enzyme containing one heme c and four Cu atoms/subunit of Mr = 88,000, was studied by electron paramagnetic resonance (EPR) at 9.2 GHz from 6 to 80 K. In the oxidized state, low spin ferric cytochrome c was observed with gz = 3.10 and an axial Cu resonance was observed with g parallel = 2.17 and g perpendicular = 2.035. No signals were detected at g values greater than 3.10. For the Cu resonance, six hyperfine lines each were observed in the g parallel and g perpendicular regions with average separations of 45.2 and 26.2 gauss, respectively. The hyperfine components are attributed to Cu(I)-Cu(II) S = 1/2 (half-met) centers. Reduction of the enzyme with dithionite caused signals attributable to heme c and Cu to disappear; exposure of that sample to N2O for a few min caused the reappearance of the g = 3.10 component and a new Cu signal with g parallel = 2.17 and g perpendicular = 2.055 that lacked the simple hyperfine components attributed to a single species of half-met center. The enzyme lost no activity as the result of this cycle of reduction and reoxidation. EPR provided no evidence for a Cu-heme interaction. The EPR detectable Cu in the oxidized and reoxidized forms of the enzyme comprised about 23 and 20% of the total Cu, respectively, or about one spin/subunit. The enzyme offers the first example of a nitrous oxide reductase which can have two states of high activity that present very different EPR spectra of Cu. These two states may represent enzyme in two different stages of the catalytic cycle.  相似文献   

16.
Low-temperature EPR examination of rat liver microsomes from control, phenobarbital-treated, and methylcholanthrene-treated animals showed the presence of both high- and low-spin ferric cytochromes P-450 and P-448. Partially purified cytochromes P-450 (from control and phenobarbital-treated rats) and P-448 (from methylcholanthrene-treated rats) were also examined with EPR. In all cases, both high- and low-spin ferric forms of cytochromes P-450 and P-448 could be observed and were found to be essentially identical compared to the microsomal preparations. However, the level of high-spin species in the soluble P-448 preparation from methylcholanthrene-treated animals was less than could be observed in the liver microsomes from the same animals. The addition of substrates increased the concentration of the high-spin form in the soluble preparations obtained from drug-treated animals. Thus, cytochromes P-450 and P-448 exist as mixtures of high- and low-spin forms. It is concluded that the substrate specificity of these cytochromes is not predetermined by the spin state of the hemoprotein. In all liver microsomal and soluble preparations, the low-spin ferric form of the hemoprotein consisted of more than a single species as determined from the EPR examinations. Each of these species upon reduction and the addition of CO yielded an identical optical spectrum. In all cases, for the ferric protein, a mercaptide sulfur is believed to be a heme ligand while the other heme ligand is variable.  相似文献   

17.
Human lactotransferrin is able to bind two vanadyl(IV) ions in specific metal-binding sites. The EPR signals of the two vanadyl bound ions, however, appear as one. This result suggests that the environments of the binding sites of human lactotransferrin are similar. The binding activity is promoted to pH 4 using carbonate or bicarbonate as synergistic anion. This unusual stability of the anion-binding site, which is destroyed below pH 6 for other transferrins, can explain in part the great stability of the metallic complexes of human lactotransferrin. However, the different sensitivities of the two metal-binding sites towards protonation permit the preparation of mixed vanadyl(IV), iron(III) complexes with VO2+ bound either on the N-terminal (acid-labile or B site) or on the C-terminal (acid-stable or A site) site. Analysis of the spectra of such mixed complexes shows the presence of a third nonspecific VO2+-binding site termed A'. The nonspecific A' site seems to be located on the outer surface of the protein close to the C-terminal site.  相似文献   

18.
The new complex [VO(chrysin)2EtOH]2 (VOchrys) has been synthesized and thoroughly characterized. Fourier transform IR, UV–vis, diffuse reflectance, and EPR spectroscopies as well as elemental analysis and thermal measurements were performed. In solution, different species could be detected by EPR spectroscopy as a function of the ligand-to-metal ratio. The stoichiometry of the chelate complex formed at pH 5 was also determined by spectrophotometric titrations. Since flavonoids are natural antioxidant compounds, the antioxidant capacity of chrysin and its vanadyl(IV) complex was investigated using different radicals. Chrysin and its complex were not able to diminish the level of superoxide and 1,1-diphenyl-2-picrylhydrazyl radicals to a great extent. In contrast, they were strong scavengers for 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonic acid diammonium salt radical cations and OH· radicals with a greater potency for VOchrys. Taking into account their selective antioxidant properties, we investigated the bioactivity of these compounds in two osteoblast-like cells in culture. Chrysin and VOchrys caused an inhibition of cell proliferation in MC3T3E1 normal osteoblasts and UMR106 tumor cells in a dose-response manner, with a greater effect in the latter cell line. The generation of reactive oxygen species (ROS) was evaluated in both cell lines and a correlation could be established between the antiproliferative effects of chrysin and the increase in the ROS levels. The complex did not generate types of ROS that can be detected by the dihydrorhodamine 123 technique so the antiproliferative effect may be attributed to the formation of other radicals such as superoxide, which is not detected by this probe. The morphological alterations were in agreement with these changes.  相似文献   

19.
The DNA binding orientation and dynamic behavior of Cu(II) complexes of 1,4,7-triazacyclononane ([9]aneN(3)), 1, and an acridine conjugate, 2, were investigated by DNA fiber EPR (EPR=electron paramagnetic resonance) spectroscopy. Crystal and molecular structure of 2 were determined by X-ray diffraction. It has been shown that 1 binds to DNA in two different modes at room temperature; one species is rapidly rotating and the other is immobilized randomly on the DNA. The introduction of acridine to [9]aneN(3) fixed the [Cu([9]aneN(3))](2+) moiety of 2 in two different environments on the DNA: the g(mid R:mid R:) axis of one species (g( parallel)=2.26) is aligned perpendicularly to the DNA fiber axis whereas that of the other (g( parallel)=2.24) aligns<90 degrees with the DNA fiber axis. The different DNA binding structures of 1 and 2 are reflected also in their different efficiencies of DNA cleavage; 2 was found to be more effective both in oxidative and hydrolytic cleavage reactions.  相似文献   

20.
Structures, chemical properties, and in vitro insulinomimetic activities of new vanadyl [oxovanadium(IV), VO(2+)] complexes with five tripodal ligands containing an imidazole functionality were examined. The ligands, N-(carboxymethyl)- N-(4-imidazolylmethyl)amino acids, contain glycine, ( S)- and ( R)-alanine, and ( S)- and ( R)-leucine residues. The molecular structures of the latter four alanine- and leucine-containing complexes were determined by X-ray analysis. The coordination geometry around each vanadium center was octahedral, where an imino nitrogen occupied the apical site and two carboxylate oxygens, an imidazole nitrogen, and a water molecule coordinated in the equatorial plane. The spectroscopic properties of the complexes were characterized by means of IR, electronic absorption, and CD spectra. Acid dissociation constants (p K(a)) and protonation sites of the ligands were determined by a combination of potentiometric titrations and (1)H NMR spectra. The potentiometric study demonstrated that stability constants (log beta) were not so different among the present complexes (14.0-14.9) and a species of molecular complex with a 1:1 metal:ligand ratio existed predominantly at physiological pH 7.4. EPR parameters indicated that the species at pH 7.4 had an octahedral structure similar to the complex in the solid state. On the other hand, an EPR study in phosphate buffer (pH 7.4) suggested that inorganic phosphate coordinated to the vanadium center instead of the imidazole group in the presence of excess phosphate ion. Cyclic voltammograms in the phosphate buffer showed chemically reversible oxidation waves, whereas irreversible oxidation waves were observed in non-coordinating HEPES buffer. Moreover, the oxidation potential of each complex in phosphate buffer was more positive than that in HEPES buffer. Partition coefficients of the present complexes in a n-octanol/saline system were very low, probably due to hydrophilicity of the imidazole group. The in vitro insulinomimetic activities were estimated on the basis of the ability of the complexes to inhibit epinephrine-stimulated free fatty acid release from isolated rat adipocytes. The achiral glycine-derivative complex exhibited the highest insulinomimetic activity, which was higher than that of VOSO(4) as a positive control. Putting our previous observations together, it was found that the vanadyl complexes with tetradentate amino acid derivatives having no alkyl side chain tend to have high in vitro insulinomimetic activity.  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