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.
The equilibria in the system VO2+ +oxidized glutathione in aqueous solution have been studied in the pH range 2-11 by a combination of pH potentiometry and spectroscopy (EPR, visible absorption and circular dichroism). The results of the various methods are self-consistent and the equilibrium model includes the species MLH4, MLH3, MLH2, MLH, ML, MLH(-1), MLH(-2) and several hydrolysis products (where H4L denotes oxidized glutathione); individual formation constants and spectra are given. Plausible structures for each stoichiometry are discussed.  相似文献   

2.
The complexation of aluminium(III) with glycerol-1-phosphate (G1P) and glycerol-2-phosphate (G2P) in aqueous solutions has been studied as a function of pH, by pH-potentiometry, 31P NMR spectroscopy and ESI mass spectrometry. Various mononuclear complexes (MLH(2)(3+), MLH(2+), ML(+), ML(2)H, ML(2)(-)) and polynuclear species (M(3)L(3)H(-1)(2+), M(3)L(2)H(-n)((n-5)-) with n=5, 6, 7, M(2)L(2)H(-1)(+) ) are formed in the system where the full protonated ligands are noted LH(2). NMR experiments clearly show that G1P and G2P already interact with Al(III) at pH 1. The potentiometric results are confirmed by ESI measurements and 31P NMR studies. No metal ion-induced deprotonation and coordination of the alcoholic-OH functions seem to occur during the complexation. The situation is very different for the glycerophosphorylcholine ligand (GPC identical with LH). Only the complex ML(3+) is formed in aqueous solution with a relatively low formation constant (K=5 at 37 degrees C). This species is clearly identified in 31P and 27Al NMR spectra. The complexation study as a function of the temperature allowed us to determine the thermodynamic parameters of the complex formation. The complexation is not governed by the reaction enthalpy that is found to be positive but by the entropy that is largely positive.  相似文献   

3.
The interaction of aluminium(III) with glucose-6-phosphate (GP: LH2) in aqueous solutions has been studied from pH 1 to pH 8, by pH-potentiometry and multinuclear (31P, 27Al, 13C) NMR spectroscopy. Various mononuclear species (MLH2, MLH, ML, ML2H, ML2 and MLH(-3)) and dinuclear complexes M2L2H-n (n=1-4) are formed in the system. NMR clearly indicates that GP is already bound to Al(III) at pH 1. The potentiometric speciation results are confirmed and completed by spectroscopic experiments. Many peaks are observed in the 31P NMR spectra suggesting the formation of isomeric species. An attempt to assign the signals to the corresponding complexes is made, allowing a discussion about their structure. Interestingly enough no metal ion-induced deprotonation and coordination of the alcoholic-OH functions have been observed.  相似文献   

4.
The obligate human pathogens Haemophilus influenzae, Neisseria gonorrhoeae, and N. meningitidis utilize a highly conserved, three-protein ATP-binding cassette transporter (FbpABC) to shuttle free Fe(3+) from the periplasm and across the cytoplasmic membrane. The periplasmic binding protein, ferric binding protein (FbpA), is capable of transporting other trivalent cations, including Ga(3+), which, unlike Fe(3+), is not redox-active. Because of a similar size and charge as Fe(3+), Ga(3+) is widely used as a non-redox-active Fe(3+) substitute for studying metal complexation in proteins and bacterial populations. The investigations reported here elucidate the similarities and differences in FbpA sequestration of Ga(3+) and Fe(3+), focusing on metal selectivity and the resulting transport function. The thermodynamic binding constant for Ga(3+) complexed with FbpA at pH 6.5, in 50 mM 4-morpholineethanesulfonic acid, 200 mM KCl, 5 mM KH(2)PO(4) was determined by UV-difference spectroscopy as [Formula: see text] This represents a 10(5)-fold weaker binding relative to Fe(3+) at identical conditions. The unfolding/refolding behavior of Ga(3+) and Fe(3+) holo-FbpA were also studied using a matrix-assisted laser desorption/ionization time-of-flight mass spectroscopy technique, stability of unpurified proteins from rates of H/D exchange (SUPREX). This analysis indicates significant differences between Fe(3+) and Ga(3+) sequestration with regard to protein folding behavior. A series of kinetic experiments established the lability of the Ga(3+)FbpA-PO(4) assembly, and the similarities/differences of stepwise loading of Fe(3+) into apo- or Ga(3+)-loaded FbpA. These biophysical characterization data are used to interpret FbpA-mediated Ga(3+) transport and toxicity in cell culture studies.  相似文献   

5.
Equilibrium and spectroscopic (1H, 13C NMR and 119Sn M?ssbauer) studies in aqueous solution are reported for dimethyltin(IV) complexes of 2-hydroxyhippuric acid (Sal-Gly). Below pH 4, oxygen-coordinated complexes MLH and ML are formed. In the pH range 5-8.5, the species MLH(-1), predominates at any metal-to-ligand ratio. The ligand exchange of this species is slow on the NMR time scale, which allows its structural characterization by NMR spectroscopy: the coordination polyhedron around the tin atom is distorted trigonal bipyramidal, with tridentate [O-,N-,COO-] coordination of Sal-Gly, involving two equatorial methyl groups. The NMR results reveal that the main cause of the distortion of the polyhedron is the large CH3-Sn-CH3 angle of 136+/-4 degrees. The presented results supplement the data available on the dimethyltin(IV)-promoted amide deprotonation of peptides, and provide further arguments for the fundamental role of the carboxylate as an anchoring group in this process.  相似文献   

6.
The equilibria of the complexation processes of V(3+) with L-alanine and L-aspartic acid in aqueous solution over a wide pH range (2-10) were studied by potentiometric and spectroscopic (UV-Vis, CD) methods. The results show that alanine forms complexes with V(3+) in the metal ion concentration range and at the ligand-to-metal ratios investigated, giving mononuclear species only. In ML(2) species, which dominate in the range pH 4-8, alanine acts as a bidendate ligand through O and N atoms. The complexation processes of V(3+) with aspartic acid are more complicated. In acidic solution (up to pH approximately 4) they are similar to those for alanine. In the higher pH region, however, there are complicated equilibria among mono- and various dinuclear species. These dinuclear species consist of carboxylic or mu-oxo bridges and differ from each other by the number of coordinated ligands and OH(-) groups. The solid phase of the V(III) complex with aspartic acid could be isolated from nonaqueous solution only. Spectroscopic (UV-Vis-IR) measurements and magnetic susceptibility data confirm the coordination of vanadium(III) by two carboxylic groups. Both V(III)-L-aspartic acid and V(III)-L-alanine complexes have a significant apoptotic effect on Hepatoma Morris 5123 cells.  相似文献   

7.
The substitution reactions of [PtCl(bpma)]+, [PtCl(gly-met-S,N,N)], [Pt(bpma)(H(2)O)](2+) and [Pt(gly-met-S,N,N)(H(2)O)](+) [where bpma is bis(2-pyridylmethyl)amine and gly-met-S,N,N is glycylmethionine] with L-methionine, glutathione and guanosine 5'-monophosphate (5'-GMP) were studied in aqueous solutions in 0.10 M NaClO(4) under pseudo-first-order conditions as a function of concentration and temperature using UV-vis spectrophotometry. The reactions of the chloro complexes were followed in the presence of 10 mM NaCl and at pH approximately 5, whereas the reactions of the aqua complexes were studied at pH 2.5. The [PtCl(bpma)]+ complex is more reactive towards the chosen nucleophiles than [PtCl(gly-met-S,N,N)]. Also, the aqua complexes are more reactive than the corresponding chloro complexes. The activation parameters for all the reactions studied suggest an associative substitution mechanism. The reactions of [PtCl(bpma)]+ and [PtCl(gly-met-S,N,N)] with 5'-GMP were studied by using (1)H NMR spectroscopy at 298 K. The pK (a) value of the [Pt(gly-met-S,N,N)(H(2)O)]+ complex is 5.95. Density functional theory calculations (B3LYP/LANL2DZp) show that in all cases guanine coordination to the L(3)Pt fragment (L(3) is terpyridine, bpma, diethylenetriamine, gly-met-S,N,N) is much more favorable than the thioether-coordinated form. The calculations collectively support the experimentally observed substitution of thioethers from Pt(II) complexes by N7-GMP. This study throws more light on the mechanistic behavior of platinum antitumor complexes.  相似文献   

8.
The reaction of VCl(3) with 1,10-phenanthroline and a series of dipeptides (H(2)dip), having aliphatic as well as aromatic side chains, in methyl alcohol and in the presence of triethylamine affords vanadium(III) compounds of the general formula [V(III)(dip)(MeOH)(phen)]Cl. Aerial oxidation/hydrolysis of the vanadium(III) species gives their oxovanadium(IV) analogues of the general formula [V(IV)O(dip)(phen)]. X-ray crystallographic characterization of the [V(IV)O(dip)(phen)] compounds (where dip(2-)=Gly- L-Ala, Gly- L-Val and Gly- L-Phe) revealed that the vanadium atom possesses a severely distorted octahedral coordination and is ligated to a tridentate dip(2-) ligand at the N(amine) atom, the deprotonated N(peptide) atom and one of the O(carboxylate) atoms, as well as an oxo group and two phenanthroline nitrogen atoms. Circular dichroism characterization of the V(III)/V(IV)O(2+)-dipeptide compounds revealed a strong signal for the V(IV)O(2+) species in the visible range of the spectrum, with a characteristic pattern which may be exploited to identify the N(am), N(pep) and O(car) ligation of a peptide or a protein to V(IV)O(2+) center, and a weak Cotton effect of opposite sign to their vanadium(III) analogues. The visible spectra of the V(III)-dipeptide compounds revealed two d-d bands with high intensity, thus indicating that the covalency of the metal-donor atoms is significant, i.e. the vanadium d orbitals are significantly mixed with the ligand orbitals, and this is confirmed by the low values of their Racah B parameters. The high-intensity band of the V(IV)O(2+)-dipeptide compounds at approximately 460 nm implies also a strong covalency of the metal with the equatorial donor atoms and this was supported by the EPR spectra of these compounds. Moreover, the V(III)/V(IV)O(2+)-dipeptide complexes were characterized by EPR and IR spectroscopies as well as conductivity and magnetic susceptibility measurements.  相似文献   

9.
Selected biochemical parameters were studied in the blood of outbred, male Wistar rats which daily received to drink deionized water (Group I, control) or solutions of: sodium metavanadate (SMV; 0.100 mg V/mL)—Group II; chromium chloride (CC; 0.004 mg Cr/mL)—Group III; and SMV-CC (0.100 mg V and 0.004 mg Cr/mL)—Group IV for a 12-week period. The diet and fluid intake, body weight gain, and food efficiency ratio (FER) diminished significantly in the rats of Groups II and IV, compared with Groups I and III. The plasma total antioxidant status (TAS) as well as the MDA and the l-ascorbic acid level in the erythrocytes (RBCs) remained unchanged in all the groups, whereas the plasma l-ascorbic acid concentration decreased markedly in Group II, compared with Group III. The activities of Cu,Zn-superoxide dismutase (Cu,Zn-SOD), catalase (CAT), cellular glutathione peroxidase (cGSH-Px), and glutathione reductase (GR) in RBCs remained unaltered in all the treated rats. However, the activity of glutathione S-transferase (GST) and the content of reduced glutathione (GSH) in RBCs decreased and increased, respectively, in Groups II, III, and IV, compared with Group I. A vanadium–chromium interaction which affected the GST activity was also found. To summarize, SMV and CC administered separately or in combination in drinking water for 12 weeks did not alter either lipid peroxidation (LPO) or the activities of Cu,Zn-SOD, CAT, cGSH-Px, and GR, which allows a conclusion that both metals in the doses ingested did not reveal their pro-oxidant potential on RBCs.  相似文献   

10.
The manganese meso-dimethylimidazolium porphyrin complex Mn(III)[TDMImP] reacted with HOBr/OBr(-) to generate the corresponding oxo-Mn(V)[TDMImP] species. The rate of this process accelerated with increasing pH. A forward rate constant, k(for), of 1.65x10(6)M(-1)s(-1) was determined at pH 8. Under these conditions, the oxo-Mn(V) species is short-lived and is transformed into the corresponding oxo-Mn(IV) complex. A first-order rate constant, k(obs), of 0.66 s(-1) was found for this reduction process at pH 8. The mechanism of this reduction process, which was dependent on bromide ion, appeared to proceed via an intermediate Mn(III)-O-Br complex. Thus, both a fast, reversible Mn(III)-O-Br bond heterolysis and a slower homolytic pathway occur in parallel in this system. The reverse oxidation reaction between oxo-Mn(V)[TDMImP] and bromide was investigated as a function of pH. The rate of this oxo-transfer reaction (k(rev)=1.4x10(3)M(-1)s(-1) at pH 8) markedly accelerated as the pH was lowered. The observed first-order dependence of the rate on [H(+)] indicates that the reactive species responsible for bromide oxidation is a protonated oxo-hydroxo complex and the stable species present in solution at high pH is dioxo-Mn(V)[TDMImP], [O=Mn(V)=O](-). The oxo-Mn(V) species retains nearly all of the oxidative driving force of the hypohalite. The equilibrium constant K(equi)=k(for)/k(rev) for the reversible process was determined at three different pH values (K(equi)=1.15x10(3) at pH 8) allowing the measurement of the redox potentials E of oxo-Mn(V)/Mn(III) (E=1.01 V at pH 8). The redox potential for this couple was extrapolated over the entire pH scale using the Nernst relationship and compared to those of the manganese 2- and 4-meso-N-methylpyridinium porphyrin couples oxo-Mn(V)[2-TMPyP]/Mn(III)[2-TMPyP], oxo-Mn(V)[4-TMPyP]/Mn(III)[4-TMPyP], OBr(-)/Br(-) and H(2)O(2)/H(2)O. Notably, the redox potential of oxo-Mn(V)/Mn(III) for the imidazolium porphyrin approaches that of H(2)O(2)/H(2)O at low pH.  相似文献   

11.
12.
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   

13.
The uptake of the oxidation products of two oxovanadium(IV) compounds, [N,N'-ethylenebis(pyridoxylaminato)]oxovanadium(IV), V(IV)O(Rpyr(2)en), and bis-[3-hydroxy-1,2-dimethyl-4-pyridinonato]oxovanadium(IV), V(IV)O(dmpp)(2), by human erythrocytes was studied using (51)V and (1)H NMR and EPR spectroscopy. V(IV)O(Rpyr(2)en) in aerobic aqueous solution is oxidized to its V(V) counterpart and the neutral form slowly enters the cells by passive diffusion. In aerobic conditions, V(IV)O(dmpp)(2) originates V(V) complexes of 1:1 and 1:2 stoichiometry. The neutral 1:1 species is taken up by erythrocytes through passive diffusion in a temperature-dependent process; its depletion from the extracellular medium promotes the dissociation of the negatively charged 1:2 species, and the protonation of the negatively charged 1:1 species. The identity of these complexes is not maintained inside the cells, and the intracellular EPR spectra suggest N(2)O(2) or NO(3) intracellular coordinating environments. The oxidative stress induced by the oxovanadium compounds in erythrocytes was not significant at 1mM concentration, but was increased by both vanadate and oxidized V(IV)O(dmpp)(2) at 5mM. Only 1mM oxidized V(IV)O(dmpp)(2) significantly stimulated erythrocytes glucose intake (0.75+/-0.13 against 0.37+/-0.17mM/h found for the control, p<0.05).  相似文献   

14.
cis,cis-1,3,5-Triaminocyclohexane (c-TACH), its N-alkyl-derivatives (alkyl = methyl, ethyl), and trans,cis-1,3,5-triaminocyclohexane (t-TACH) were prepared, and speciation and DNA cleaving property of Cu(II) complexes of these ligands were investigated. All of the complexes efficiently promote the hydrolytic cleavage of supercoiled plasmid DNA under physiological conditions without further additives. The DNA cleavage rate (V(obs)) trend at pH values between 8 and 9 is N-Me(3) = N-Et(1) < t-TACH < c-TACH < N-Et(2) < N-Et(3). At pH 7, the trend is c-TACH < N-Et(3) = N-Et(2) < N-Et(1) < N-Me(3) < t-TACH. The cleavage rate constants at 35 degrees C, for the c-TACH complex are 3 x 10(-1) h(-1) at pH 8.1 and 2 x 10(-1) h(-1) at pH 7.0 ([DNA] = 7 microM, [Cu(II)-complex] = 105 microM). The hydrolytically active species at pH > 8 is CuL(H(2)O)(OH)(+) in which L coordinates to Cu(II) as a tridentate ligand for all complexes except for t-TACH. The hydrolytically active species at pH 7 is CuLH(H(2)O)(3)(3+) or CuLH(H(2)O)(4)(3+) in which LH coordinates as bidentate ligand. DNA-binding constants of c-TACH and t-TACH complexes are presented and the effects of N-alkyl and ammonium groups are discussed in light of the proposed reaction mechanism.  相似文献   

15.
The yeast Rhodotorula glutinis was examined for its ability to remove Pb(2+) from aqueous solution. Within 10 min of contact, Pb(2+) sorption reached nearly 80% of the total Pb(2+) sorption. The optimum initial pH value for removal of Pb(2+ )was 4.5-5.0. The percentage sorption increased steeply with the biomass concentration up to 2 g/l and thereafter remained more or less constant. Temperature in the range 15-45 degrees C did not show any significant difference in Pb(2+ )sorption by R. glutinis. The light metal ions such as Na(+), K(+), Ca(2+), and Mg(2+) did not significantly interfere with the binding. The Langmuir sorption model provided a good fit throughout the concentration range. The maximum Pb(2+ )sorption capacity q(max) and Langmuir constant b were 73.5 mg/g of biomass and 0.02 l/mg, respectively. The mechanism of Pb(2+) removal by R. glutinis involved biosorption by direct biosorptive interaction with the biomass through ion exchange and precipitation by phosphate released from the biomass.  相似文献   

16.
The reaction of [VO(CH3COO)2(phen)] (phen = 1,10-phenanthroline) with the sulfhydryl-containing pseudopeptides (scp), N-(2-mercaptopropionyl)glycine (H3mpg), N-(2-mercaptopropionyl)cysteine (H4m2pc), N-(3-mercaptopropionyl)cysteine (H4m3pc) and the dipeptides glycylglycine (H2glygly) and glycyl-L-alanine (H2glyala), in the presence of triethylamine, results in the formation of the compounds Et3NH[VO(mpg)(phen)] (1), (Et3NH)2[VO(m2pc)] (4), [(Et3NH)2[VO(m3pc) (5), [VO(glygly)(phen)] x 2CH3OH (2 x 2CH3OH) and [VO(glyala)(phen)] x CH3OH (3 x CH3OH). Evidence for the molecular connectivity in 2 x CH3OH was established by X-ray crystallography, showing the vanadium(IV) atom ligated to a tridentate glygly2- ligand at the N(amine), N(peptide) and O(carboxylato) atoms. Combination of the correlation plot of the EPR parameters gz versus Az, together with the additivity relationship supported the prediction of the equatorial donor atom sets of the V(IV)O2+ center at various pH values for the V(IV)O2+-glutathione system considered in this study. Model NMR studies (interaction of vanadium(V) with the scp H3mpg) showed that there is a possibility of vanadium(V) ligation to glutathione.  相似文献   

17.
The protonation equilibria of alanylglycylhistamine (Ala-Gly-Ha) and the complexation of this ligand with Cu(II) and Ni(II) have been studied by pH-potentiometry, 1H and 14N NMR spectroscopy, electrospray ionization mass spectrometry (ESI-MS), circular dichroism (CD), UV-Vis spectrophotometry and electron paramagnetic resonance (EPR). From pH approximately 2-12, the following complexes: MLH, MLH(-1), MLH(-2) and MLH(-3) are successively formed in aqueous solutions, the ligand under its neutral form being noted L. At physiological pH, the MLH(-2) complex is predominant. The coordination in this complex is assumed by one amino, two deprotonated peptide and one imidazole nitrogen atoms. The ESI-MS study confirmed the formation of the MLH(-1), MLH(-2) and MLH(-3) complexes. The structure of MLH(-2) was determined by single crystal X-ray analysis. CD and UV-Vis techniques allowed us to propose that the imidazole-N3 nitrogen acts as the anchor group for the coordination to the metal(II) ions rather than the amino group. At high pH values, the further deprotonation of the N-H imidazole group, leading to the formation of MLH(-3), occurs, as revealed by 1H NMR spectroscopy.  相似文献   

18.
We studied the interaction of gamma-L-glutamyl-L-cysteinyl-glycine (glutathione, GSH) with cadmium ions (Cd(2+)) by first performing classical potentiometric pH titration measurements and then turning to additional spectroscopic methods. To estimate the residual concentrations of free cadmium, we studied the competition of glutathione with a Cd(2+)-sensitive dye, either an absorbing dye (murexide) or a fluorescent one (FluoZin-1), and consistent results were obtained with the two dyes. In KCl-containing Tes, Mops, or Tris buffer at pH 7.0 to 7.1 and 37 degrees C (and at a total Cd(2+) concentration of 0.01 mM), results suggest that free cadmium concentration is halved when the concentration of glutathione is approximately 0.05 mM; this mainly reflects the combined apparent dissociation constant for the Cd(glutathione) 1:1 complex under these conditions. To identify the other complexes formed, we used far-UV spectroscopy of the ligand-to-metal charge transfer absorption bands. The Cd(glutathione)(2) 1:2 complex predominated over the 1:1 complex only at high millimolar concentrations of total glutathione and not at low submillimolar concentrations of total glutathione. The apparent conditional constants derived from these spectroscopy results made it possible to discriminate between sets of absolute constants that would otherwise have simulated the pH titration data similarly well in this complicated system. Related experiments showed that although the Cl(-) ions in our media competed (modestly) with glutathione for binding to Cd(2+), the buffers we had chosen did not bind Cd(2+) significantly under our conditions. Our experiments also revealed that Cd(2+) may be adsorbed onto quartz or glass vessel walls, reducing the accuracy of theoretical predictions of the concentrations of species in solution. Lastly, the experiments confirmed the rapid kinetics of formation and dissociation of the UV-absorbing Cd(glutathione)(2) 1:2 complexes. The methods described here may be useful for biochemists needing to determine conditional binding constants for charge transfer metal-ligand complexes under their own conditions.  相似文献   

19.
The nematocide and soil fumigant 1,2-dibromo-3-chloropropane (DBCP) is a carcinogen and a mutagen and displays target-organ toxicity to the testes and the kidney. It has been proposed that both cytochrome P-450 mediated activation and glutathione (GSH) conjugation pathways are operative in DNA damage and organotropy induced by DBCP. To determine the chemical mechanisms involved in the bioactivation of DBCP and to assess a role for an episulfonium ion intermediate, the mechanism of formation of GSH conjugate metabolites of DBCP was investigated. Five biliary GSH conjugates of DBCP were isolated from rats and identified by fast atom bombardment tandem mass spectrometry: S-(2,3-dihydroxy-propyl)glutathione (I), S-(2-hydroxypropyl)glutathione (IIA), S-(3-chloro-2-hydroxypropyl)glutathione (III), 1,3-di(S-glutathionyl)propan-2-ol (IV), and 1-(glycyl-S-cysteinyl)-3- (S-glutathionyl)propan-2-ol (V). The mechanisms of conjugate formation were addressed by assessing deuterium retention in conjugates derived from [1,1,2,3,3-2H5] DBCP (D5-DBCP). GSH conjugates I, III, IV, and V displayed quantitative retention of deuterium, an observation consistent with the formation of an episulfonium ion intermediate. GSH conjugate IIA, however, retained three atoms of deuterium, thus invoking a P-450 mechanism in its genesis. The involvement of glutathione transferase (GST) and sequential episulfonium ion intermediates in the formation of metabolites I, III, and IV was demonstrated in vitro. Upon incubation of DBCP with GST, metabolites I, III, and IV were identified by tandem mass spectrometry and were found to arise with quantitative retention of deuterium when D5-DBCP was employed as a substrate. An additional GSH conjugate, 1,2,3-tri(S-glutathionyl)propane (VI), was observed as the major metabolite in incubations of GST with DBCP. When the incubations of DBCP with GST were performed in H2(18)O, metabolite I incorporated two atoms of 18O, and metabolites III and IV incorporated one atom of 18O. The ability of GST to catalyze the formation of the four GSH conjugates observed in vivo, with quantitative retention of deuterium and incorporation of 18O from H2(18)O, may be rationalized by a mechanism invoking the initial formation of S-(2-bromo-3-chloropropyl)glutathione. Rearrangement of this unstable conjugate via several reactive episulfonium ions, with either hydrolysis by water or alkylation of GSH at various stages, would account for the pattern of metabolites and their status of isotopic enrichment observed under various incubation conditions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
Inhibition by polyvalent cations is a defining characteristic of voltage-gated proton channels. The mechanism of this inhibition was studied in rat alveolar epithelial cells using tight-seal voltage clamp techniques. Metal concentrations were corrected for measured binding to buffers. Externally applied ZnCl(2) reduced the H(+) current, shifted the voltage-activation curve toward positive potentials, and slowed the turn-on of H(+) current upon depolarization more than could be accounted for by a simple voltage shift, with minimal effects on the closing rate. The effects of Zn(2+) were inconsistent with classical voltage-dependent block in which Zn(2+) binds within the membrane voltage field. Instead, Zn(2+) binds to superficial sites on the channel and modulates gating. The effects of extracellular Zn(2+) were strongly pH(o) dependent but were insensitive to pH(i), suggesting that protons and Zn(2+) compete for external sites on H(+) channels. The apparent potency of Zn(2+) in slowing activation was approximately 10x greater at pH(o) 7 than at pH(o) 6, and approximately 100x greater at pH(o) 6 than at pH(o) 5. The pH(o) dependence suggests that Zn(2+), not ZnOH(+), is the active species. Evidently, the Zn(2+) receptor is formed by multiple groups, protonation of any of which inhibits Zn(2+) binding. The external receptor bound H(+) and Zn(2+) with pK(a) 6.2-6.6 and pK(M) 6.5, as described by several models. Zn(2+) effects on the proton chord conductance-voltage (g(H)-V) relationship indicated higher affinities, pK(a) 7 and pK(M) 8. CdCl(2) had similar effects as ZnCl(2) and competed with H(+), but had lower affinity. Zn(2+) applied internally via the pipette solution or to inside-out patches had comparatively small effects, but at high concentrations reduced H(+) currents and slowed channel closing. Thus, external and internal zinc-binding sites are different. The external Zn(2+) receptor may be the same modulatory protonation site(s) at which pH(o) regulates H(+) channel gating.  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