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.
Iron uptake by transferrin from triacetohydroxamatoFe(III) (Fe(AHA)3) in the presence of bicarbonate has been investigated between pH 7 and 8.2. The protein transits from the opened apo- to the closed holoform by several steps with the accumulation of at least three kinetic intermediates. All these steps are accompanied by proton losses, probably occurring from the protein ligands and the side-chains involved in the interdomain H-bonding nets. The minor bihydroxamatoFe(III) species Fe(AHA)2 exchanges its iron with the C-site of apotransferrin in interaction with bicarbonate without detectable formation of any intermediate protein-iron-ligand mixed complex; direct second-order rate constant k1=4.15(+/-0.05)x10(7) M(-1) s(-1). The kinetic product loses a single proton and undergoes a modification in its conformation followed by the loss of two or three protons; first-order rate constant k2=3.25(+/-0.15) s(-1). This induces a new modification in the conformation; first-order rate constant k3=5.90(+/-0.30)x10(-2) s(-1). This new modification in conformation rate controls iron uptake by the N-site of the protein and is followed by a single proton loss; K3a=6.80 nM. Finally, the holoprotein or the monoferric transferrin in its thermodynamic equilibrated state is produced by a last modification in the conformation occurring in about 4000 seconds. But for the Fe(AHA)3 dissociation and the involvement of Fe(AHA)2 in the first step of iron uptake, this mechanism is identical to that reported for iron uptake from FeNAc3. This implies that the exchange of iron between a chelate and serum-transferrin occurs by a single general mechanism. The nature of the iron-providing chelate is only important for the first kinetic step of the exchange, which can be slowed to such an extent that it rate limits the exchange of iron.  相似文献   

2.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

3.
The kinetics and thermodynamics of Bi(III) exchange between bismuth mononitrilotriacetate (BiL) and human serum transferrin as well as those of the interaction between bismuth-loaded transferrin and transferrin receptor 1 (TFR) were investigated at pH 7.4-8.9. Bismuth is rapidly exchanged between BiL and the C-site of human serum apotransferrin in interaction with bicarbonate to yield an intermediate complex with an effective equilibrium constant K(1) of 6 +/- 4, a direct second-order rate constant k(1) of (2.45 +/- 0.20) x 10(5) M(-1) s(-1), and a reverse second-order rate constant k(-1) of (1.5 +/- 0.5) x 10(6) M(-1) s(-1). The intermediate complex loses a single proton with a proton dissociation constant K(1a) of 2.4 +/- 1 nM to yield a first kinetic product. This product then undergoes a modification in its conformation followed by two proton losses with a first-order rate constant k(2) = 25 +/- 1.5 s(-1) to produce a second kinetic intermediate, which in turn undergoes a last modification in the conformation to yield the bismuth-saturated transferrin in its final state. This last process rate-controls Bi(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau(3)(-1) of (3 +/- 1) x 10(-2) s(-1). The mechanism of bismuth uptake differs from that of iron and probably does not involve the same transition in conformation from open to closed upon iron uptake. The interaction of bismuth-loaded transferrin with TFR occurs in a single very fast kinetic step with a dissociation constant K(d) of 4 +/- 0.4 microM, a second-order rate constant k(d) of (2.2 +/- 1.5) x 10(8) M(-1) s(-1), and a first-order rate constant k(-d) of 900 +/- 400 s(-1). This mechanism is different from that observed with the ferric holotransferrin and implies that the interaction between TFR and bismuth-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of bismuth incorporation by the transferrin receptor-mediated iron acquisition pathway is discussed.  相似文献   

4.
Slc26a2 is a ubiquitously expressed SO(4)(2-) transporter with high expression levels in cartilage and several epithelia. Mutations in SLC26A2 are associated with diastrophic dysplasia. The mechanism by which Slc26a2 transports SO(4)(2-) and the ion gradients that mediate SO(4)(2-) uptake are poorly understood. We report here that Slc26a2 functions as an SO(4)(2-)/2OH(-), SO(4)(2-)/2Cl(-), and SO(4)(2-)/OH(-)/Cl(-) exchanger, depending on the Cl(-) and OH(-) gradients. At inward Cl(-) and outward pH gradients (high Cl(-)(o) and low pH(o)) Slc26a2 functions primarily as an SO(4)(2-)(o)/2OH(-)(i) exchanger. At low Cl(-)(o) and high pH(o) Slc26a2 functions increasingly as an SO(4)(2-)(o)/2Cl(-)(i) exchanger. The reverse is observed for SO(4)(2-)(i)/2OH(-)(o) and SO(4)(2-)(i)/2Cl(-)(o) exchange. Slc26a2 also exchanges Cl(-) for I(-), Br(-), and NO(3)(-) and Cl(-)(o) competes with SO(4)(2-) on the transport site. Interestingly, Slc26a2 is regulated by an extracellular anion site, required to activate SO(4)(2-)(i)/2OH(-)(o) exchange. Slc26a2 can transport oxalate in exchange for OH(-) and/or Cl(-) with properties similar to SO(4)(2-) transport. Modeling of the Slc26a2 transmembrane domain (TMD) structure identified a conserved extracellular sequence (367)GFXXP(371) between TMD7 and TMD8 close to the conserved Glu(417) in the permeation pathway. Mutation of Glu(417) eliminated transport by Slc26a2, whereas mutation of Phe(368) increased the affinity for SO(4)(2-)(o) 8-fold while reducing the affinity for Cl(-)(o) 2 fold, but without affecting regulation by Cl(-)(o). These findings clarify the mechanism of net SO(4)(2-) transport and describe a novel regulation of Slc26a2 by an extracellular anion binding site and should help in further understanding aberrant SLC26A2 function in diastrophic dysplasia.  相似文献   

5.
In growing neurite of neuronal cells, it is suggested that α/β-tubulin heterodimers assemble to form microtubule, and assembly of microtubule promotes neurite elongation. On the other hand, recent studies reveal importance of intracellular Cl(-) in regulation of various cellular functions such as cell cycle progression, differentiation, cell migration, and elongation of neurite in neuronal cells. In this study, we investigated effects of Cl(-) on in vitro tubulin polymerization. We found that efficiency of in vitro tubulin polymerization (the number of microtubule) was higher (3 to 5-fold) in Cl(-)-containing solutions than that in Cl(-)-free solutions containing Br(-) or NO(3)(-). On the other hand, GTPase activity of tubulin was lower (2/3-fold) in Cl(-)-containing solutions than that in Cl(-)-free solutions containing Br(-) or NO(3)(-). Efficiency of in vitro tubulin polymerization in solutions containing a non-hydrolyzable analogue of GTP (GpCpp) instead of GTP was much higher than that in the presence of GTP. Effects of replacement of GTP with GpCpp on in vitro tubulin polymerization was weaker in Cl(-) solutions (10-fold increases) than that in Br(-) or NO(3)(-) solutions (20-fold increases), although the efficiency of in vitro tubulin polymerization in Cl(-) solutions containing GpCpp was still higher than that in Br(-) or NO(3)(-) solutions containing GpCpp. Our results suggest that a part of stimulatory effects of Cl(-) on in vitro tubulin polymerization is mediated via an inhibitory effect on GTPase activity of tubulin, although Cl(-) would also regulate in vitro tubulin polymerization by factors other than an inhibitory effect on GTPase activity.  相似文献   

6.
On the basis of the anion content of in vitro-cultured Arabidopsis plantlets, we explored the selectivity of the voltage-dependent anion channel of the plasma membrane of hypocotyl cells. In the whole-cell configuration, substitution of cytosolic Cl(-) by different anions led to the following sequence of relative permeabilities: NO(3)(-) (2.6) >/= SO(4)(2-) (2.0) > Cl(-) (1.0) > HCO(3)(-) (0.8) > malate(2-) (0.03). Large whole-cell currents were measured for NO(3)(-) and SO(4)(2-), about five to six times higher than the equivalent Cl(-) currents. Since SO(4)(2-) is usually considered to be a weakly permeant or non-permeant ion, the components of the large whole-cell current were explored in more detail. Aside from its permeation through the channel with a unitary conductance, about two-thirds that of Cl(-), SO(4)(2-) had a regulatory effect on channel activity by preventing the run-down of the anion current both in the whole-cell and the outside-out configuration, increasing markedly the whole-cell current. The fact that the voltage-dependent plasma membrane anion channel of hypocotyl cells can mediate large NO(3)(-) and SO(4)(2-) currents and is regulated by nucleotides favors the idea that this anion channel can contribute to the cellular homeostasis of important metabolized anions.  相似文献   

7.
The inhibitory effects of anions, such as N(3)(-), NO(2)(-), BO(4)(3-), SCN(-), CH(3)COO(-), SO(4)(2-), ClO(4)(-), H(2)PO(4)(-), CN(-), I(-), Br(-), Cl(-) and F(-), on the hydrolysis of L-arginine (L-Arg) by rat liver arginase (RLA) have been studied. From all these anions, only F(-) exhibited a clear inhibitory effect at the mM level. Inhibition of RLA by F(-) is reversible and uncompetitive towards L-Arg binding with a K(i) value of 1.3+/-0.5 mM at pH 7.4. This effect is dependent on pH as the IC(50) value of F(-) towards RLA increases from 1.2 to 19 mM when increasing the pH from 7 to 10. Another specific inhibitor of RLA, N(omega)-hydroxy-L-nor-arginine (nor-NOHA), that has been recently shown to bind to RLA as a bridging ligand of its (Mn(II))(2) cluster, exhibits some similarities with F(-) in its inhibitory effects (identical pH dependence). It is thus tempting to propose that the inhibitory effects of F(-) could be due to its binding as a bridging ligand of the RLA (Mn(II))(2) cluster. However, further studies are required to determine the modes of interaction of F(-) with RLA.  相似文献   

8.
We report the determination in cell-free assays of the mechanism of iron release from the N-lobe and C-lobe of human serum transferrin in interaction with intact transferrin receptor 1 at 4.3< or =pH< or =6.5. Iron is first released from the N-lobe in the tens of milliseconds range and then from the C-lobe in the hundreds of seconds range. In both cases, iron loss is rate-controlled by slow proton transfers, rate constant for the N-lobe k(1)=1.20(+/-0.05)x10(6)M(-1)s(-1) and for the C-lobe k(2)=1.6(+/-0.1)x10(3)M(-1)s(-1). This iron loss is subsequent to a fast proton-driven decarbonation and is followed by two proton gains, (pK(1a))/2=5.28 per proton for the N-lobe and (pK(2a))/2=5.10 per proton for the C-lobe. Under similar experimental conditions, iron loss is about 17-fold faster from the N-lobe and is at least 200-fold faster from the C-lobe when compared to holotransferrin in the absence of receptor 1. After iron release, the apotransferrin-receptor adduct undergoes a slow partial dissociation controlled by a change in the conformation of the receptor; rate constant k(3)=1.7(+/-0.1)x10(-3)s(-1). At endosomic pH, the final equilibrated state is attained in about 1000 s, after which the free apotransferrin, two prototropic species of the acidic form of the receptor and apotransferrin interacting with the receptor coexist simultaneously. However, since recycling of the vesicle containing the receptor to the cell surface takes a few minutes, the major part of transferrin will still be forwarded to the biological fluid in the form of the apotransferrin-receptor protein-protein adduct.  相似文献   

9.
Members of the SLC26 transporter family play an essential role in several epithelial functions, as revealed by diseases associated with mutations in members of the family. Several members were shown to function as Cl(-) and HCO(3)(-) transporters that likely play an important role in epithelial Cl(-) absorption and HCO(3)(-) secretion. However, the mechanism of most transporters is not well understood. SLC26A7 is a member of the SLC26 transporter family reported to be expressed in the basolateral membrane of the cortical collecting duct and parietal cells and functions as a coupled Cl(-)/HCO(3)(-) exchanger. In the present work we examined the transport properties of SLC26A7 to determine its transport characteristics and electrogenicity. We found that when expressed in Xenopus oocytes or HEK293 cells SLC26A7 functions as a pH(i)-regulated Cl(-) channel with minimal OH(-)/HCO(3)(-) permeability. Expression of SLC26A7 in oocytes or HEK293 cells generated a Cl(-) current with linear I/V and an instantaneous current that was voltage- and time-independent. Based on measurement of reversal potential the selectivity of SLC26A7 is NO(3)(-)>Cl(-)=Br(-)=I(-)>SO(4)(2-)=Glu(-), although I(-) partially inhibited the current. Incubating the cells with HCO(3)(-) or butyrate acidified the cytosol and increased the selectivity of SLC26A7 for Cl(-). Measurement of membrane potential and pH(i) showed minimal OH(-) and HCO(3)(-) transport by SLC26A7 when the cells were incubated in Cl(-)-containing or Cl(-)-free media. The activity of SLC26A7 was inhibited by all inhibitors of anion transporters tested, 4,4'-diisothiocyanostilbene-2,2'-disulfonic acid, diphenylamine-2-carboxylic acid, and glybenclamide. These findings reveal that SLC26A7 functions as a unique Cl(-) channel that is regulated by intracellular H(+).  相似文献   

10.
The SLC4A1/AE1 gene encodes the electroneutral Cl(-)/HCO(3)(-) exchanger of erythrocytes and renal type A intercalated cells. AE1 mutations cause familial spherocytic and stomatocytic anemias, ovalocytosis, and distal renal tubular acidosis. The mutant mouse Ae1 polypeptide E699Q expressed in Xenopus oocytes cannot mediate Cl(-)/HCO(3)(-) exchange or (36)Cl(-) efflux but exhibits enhanced dual sulfate efflux mechanisms: electroneutral exchange of intracellular sulfate for extracellular sulfate (SO(4)(2-)(i)/SO(4)(2-)(o) exchange), and electrogenic exchange of intracellular sulfate for extracellular chloride (SO(4)(2-)(i)/Cl(-)(o) exchange). Whereas wild-type AE1 mediates 1:1 H(+)/SO(4)(2-) cotransport in exchange for either Cl(-) or for the H(+)/SO(4)(2-) ion pair, mutant Ae1 E699Q transports sulfate without cotransport of protons, similar to human erythrocyte AE1 in which the corresponding E681 carboxylate has been chemically converted to the alcohol (hAE1 E681OH). We now show that in contrast to the normal cis-stimulation by protons of wild-type AE1-mediated SO(4)(2-) transport, both SO(4)(2-)(i)/Cl(-)(o) exchange and SO(4)(2-)(i)/SO(4)(2-)(o) exchange mediated by mutant Ae1 E699Q are inhibited by acidic pH(o) and activated by alkaline pH(o). hAE1 E681OH displays a similarly altered pH(o) dependence of SO(4)(2-)(i)/Cl(-)(o) exchange. Elevated [SO(4)(2-)](i) increases the K(1/2) of Ae1 E699Q for both extracellular Cl(-) and SO(4)(2-), while reducing inhibition of both exchange mechanisms by acid pH(o). The E699Q mutation also leads to increased potency of self-inhibition by extracellular SO(4)(2-). Study of the Ae1 E699Q mutation has revealed the existence of a novel pH-regulatory site of the Ae1 polypeptide and should continue to provide valuable paths toward understanding substrate selectivity and self-inhibition in SLC4 anion transporters.  相似文献   

11.
The time course of the relaxation effect induced by a single dose (3 x 10(-6) mol/L) of trans-[Ru(NH3)4L(NO)]3+ (L=nic, 4-pic, py, imN, P(OEt)3, SO(3)(2-), NH3, and pz) species and sodium nitroprusside (4 x 10(-9) mol/L) was studied in aortic rings without endothelium and pre-contracted with noradrenaline (1 x 10(-6) mol/L). All the compounds induced a relaxing effect in the aortic rings, but the intensity and time of relaxation were different. Only the species where L=py, 4-pic, and P(OEt)3 were able to induce 100% (99-100%) of the relaxing effect during the assay. trans-[Ru(NH3)4(L)(NO)]3+ (L=SO(3)(2-) and NH3) showed the lowest relaxing effect (36 and 37%, respectively) when compared with the other compounds. Relationship was observed between the time corresponding to half of the maximum relaxation intensity observed and, respectively, k-NO, E0'[Ru(NO)]3+/[Ru(NO)]2+ in trans-[Ru(NH3)4(L)(NO)]3+ species and E0'Ru(III)/Ru(II) in trans-[Ru(NH3)4(L)(H2O)]3+ ions. These relationships strongly suggested that the NO liberation from the reduced nitrosyl complexes was responsible for the observed relaxation.  相似文献   

12.
Effects of nitrite and ammonium on methane-dependent denitrification   总被引:1,自引:0,他引:1  
For effective application of methane-dependent denitrification (MDD) in the treatment of wastewater containing NO(2)(-) or NH(4)(+), the effect of these inorganic nitrogen compounds on MDD activity needs to be clarified. The MDD activity of sludge acclimatized with CH(4) and O(2) was determined with mineral media of different nitrogen-compound compositions in the presence of 0.21 atm CH(4) and 0.20 atm O(2). Incubations with media containing only NO(2)(-) or two of the three inorganic nitrogen compounds (NO(3)(-)+NO(2)(-), NO(2)(-)+NH(4)(+) or NH(4)(+)+NO(3)(-)) resulted in MDD activity equal to or higher than that with media containing only NO(3)(-). However, there was no MDD activity in media containing NO(2)(-) at 10 degrees C, probably because of serious inhibition of NO(2)(-) on methane oxidation. MDD occurred in media containing only NH(4)(+), although the total nitrogen removal efficiency was very low. These results show that NO(2)(-) and NH(4)(+), in the presence of NO(x)(-), do not inhibit but rather promote MDD. Consequently, NH(4)(+) does not need to be completely oxidized to NO(3)(-) in the nitrification reactor before MDD. However, under psychrophilic conditions, NO(2)(-) seriously inhibited MDD. Therefore, the nitrification reactor must not discharge effluent containing NO(2)(-) under psychrophilic conditions.  相似文献   

13.
Eels are unique in that they maintain lower plasma SO(4)(2-) concentration in SO(4)(2-)-rich (~30 mM) seawater (SW) than in SO(4)(2-)-poor (<0.3 mM) freshwater (FW), showing drastic changes in SO(4)(2-) regulation between FW and SW. We previously showed that the expression of renal SO(4)(2-) transporter genes, FW-specific Slc13a1 and SW-specific Slc26a6a, changes profoundly after transfer of FW eels to SW, which results in the decrease in plasma SO(4)(2-) concentration after 3 days in SW. In this study, we attempted to identify the environmental factor(s) that trigger the switching of SO(4)(2-) regulation using changes in plasma and urine SO(4)(2-) concentrations and expression of the transporter genes as markers. Transfer of FW eels to 30 mM SO(4)(2-) or transfer of SW eels to SO(4)(2-)-free SW did not change the SO(4)(2-) regulation. Major divalent cations in SW, Mg(2+) (50 mM) and Ca(2+) (10 mM), were also ineffective, but 50 mM NaCl was effective for switching the SO(4)(2-) regulation. Further analyses using choline-Cl and Na-gluconate showed that Cl(-) is a primary factor and Na(+) is permissive for the Cl(-) effect. Since plasma SO(4)(2-) and Cl(-) concentrations were inversely correlated, we injected various solutions into the blood and found that Cl(-) alone triggered the switching from FW to SW-type regulation. Furthermore, the inhibitor of Na-Cl cotransporter (NCC) added to media significantly impaired the expression of SW-specific Slc26a6a in 150 mM NaCl. In summary, it appears that Cl(-) ions in SW are taken up into the circulation via the NCC together with Na(+), and the resultant increase in plasma Cl(-) concentration enhances SO(4)(2-) excretion by the kidney through downregulation of absorptive Slc13a1 and upregulation of excretory Slc26a6a, resulting in low plasma SO(4)(2-) concentration in SW.  相似文献   

14.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

15.
Transferrin's mechanism of interaction with receptor 1   总被引:1,自引:0,他引:1  
The kinetics and thermodynamics of the interactions of transferrin receptor 1 with holotransferrin and apotransferrin in neutral and mildly acidic media are investigated at 37 degrees C in the presence of CHAPS micelles. Receptor 1 interacts with CHAPS in a very fast kinetic step (<1 micros). This is followed in neutral media by the interaction with holotransferrin which occurs in two steps after receptor deprotonation, with a proton dissociation constant (K(1a)) of 10.0 +/- 1.5 nM. The first step is detected by the T-jump technique and is associated with a molecular interaction between the receptor and holotransferrin. It occurs with a first-order rate constant (k(-1)) of (1.6 +/- 0.2) x 10(4) s(-1), a second-order rate constant (k(1)) of (3.20 +/- 0.2) x 10(10) M(-1) s(-1), and a dissociation constant (K(1)) of 0.50 +/- 0.07 microM. This step is followed by a slow change in the conformation with a relaxation time (tau(2)) of 3400 +/- 400 s and an equilibrium constant (K(2)) of (4.6 +/- 1.0) x 10(-3) with an overall affinity of the receptor for holotransferrin [(K'1)(-1)] of (4.35 +/- 0.60) x 10(8) M(-1). Apotransferrin does not interact with receptor 1 in neutral media, between pH 4.9 and 6, it interacts with the receptor in two steps after a receptor deprotonation (K(2a) = 2.30 +/- 0.3 microM). The first step occurs in the range of 1000-3000 s. It is ascribed to a slow change in the conformation which rate-controls a fast interaction between apotransferrin and receptor 1 with an overall affinity constant [(K(3))(-1)] of (2.80 +/- 0.30) x 10(7) M(-1). These results imply that receptor 1 probably exists in at least two forms, the neutral species which interacts with holotransferrin and not with apotransferrin and the acidic species which interacts with apotransferrin. At first, the interaction of the neutral receptor with holotransferrin is extremely fast. It is followed by the slow change in conformation, which leads to an important stabilization of the thermodynamic structure. In the acidic media of the endosome, the interaction of apotransferrin with the acidic receptor is sufficiently strong and rate-controlled by a very slow change in conformation which allows recycling back to the plasma membrane.  相似文献   

16.
The push effect of anionic axial ligands of high-valent iron(IV)-oxo porphyrin pi-cation radicals, (Porp)(+.)Fe(IV)(O)(X) (X=OH(-), AcO(-), Cl(-), and CF(3)SO(3)(-)), in alkane hydroxylation is investigated by B3LYP DFT calculations. The electron-donating ability of anionic axial ligands influences the activation energy for the alkane hydroxylation by the iron(IV)-oxo intermediates and the Fe-O bond distance of the iron-oxo species in transition state.  相似文献   

17.
The kinetics and mechanisms of the reactions of a number of pyrogallol-based ligands with iron(III) have been investigated in aqueous solution at 25 degrees C and ionic strength 0.5 M NaClO(4). Mechanisms have been proposed which account satisfactorily for the kinetic data. These are generally consistent with a mechanism in which the 1:1 complex that is formed initially when the metal reacts with the ligand subsequently decays through an electron transfer reaction. There was also some evidence for the formation of a 1:2 ligand-to-metal complex at higher pH values. The kinetics of complex formation were investigated with either the ligand or metal in pseudo-first-order excess. Rate constants for k(1) of 2.83(+/-0.09)x10(3), 1.75(+/-0.045)x10(3) and 3300(+/-200) M(-1) s(-1) and k(-1) of 20(+/-6.0), 35(+/-13) and 25+/-7.6 M(-1) s(-1) have been evaluated for the reaction of Fe(OH)(2+) with gallic acid, gallic acid methyl ester and catechin, respectively. The stability constant of each [Fe(L)](+) complex has been calculated from the kinetic data. The iron(III) assisted decomposition of the initial iron(III) complex formed was investigated. Analysis of the kinetic data yielded both the equilibrium constants for protonation of the iron(III) complexes initially formed together with the rate constants for the intramolecular electron transfers for gallic acid and gallic acid methyl ester. All of the suggested mechanisms and calculated rate constants are supported by calculations carried out using global analysis of time-dependent spectra.  相似文献   

18.
The uptake of Cl minus by the ascomycetes Neocosmospora vasinfecta was investigated. Intramycelial concentrations of more than 55 mM (890-fold the external concentration) were reached. Accumulation was as inorganic Cl minus and nystatin induced total loss of mycelial Cl minus without extensive protein loss, implying that Cl minus retention was not due to binding to macromolecules. Cl minus transport was largely unidirectional with efflux being low under all conditions. Uptake was temperature dependent (maximal Arrhenius activation energy of 18.0 kcal/mol) and was severely reduced by KCN, dicyclohexylcarbodiimide, and anaerobiosis. A comparison with the inhibition of oxygen uptake under the same conditions implied that Cl minus transport was not directly coupled to aerobic electron transport. Cl minus uptake was a saturatable function of the external Cl minus concentration, and apparent Km values of 6.4 times 10-6 M and 10-4 M were calculated. Of the anions tested, only Br minus effectively inhibited Cl minus uptake and I minus, NO3 minus, SO4 minus 2, HCO3 minus, and H2PO4 minus were without effect. Cl minus uptake did not require concomitant cation uptake.  相似文献   

19.
A voltammetric ion-channel sensing for phosphate based on gold electrodes modified with the self-assembled monolayers of a bis-thiourea receptor was developed to detect phosphate. The working principle of this voltammetric sensor conceptually mimics that of ligand gated ion-channel proteins, as to chemically stimulated changes in membrane permeability. The response to analytes is based on the change in electron transfer rate constant of the redox reaction of [Fe(CN)(6)](4-/3-) marker, before and after binding of phosphate to the receptor on the electrode surface; where the electrostatic repulsion between a phosphate-receptor complex and the marker induced the decrease in the rate constant. In a solution of pH 7.0, a high selectivity was observed for phosphate and the sensor was virtually insensitive at all to many of other anions, such as SO(4)(2-), AcO(-), NO(3)(-), and Cl(-). The sensor response was obtained with phosphate concentrations above 5.0 x 10(-4) M using cyclic voltammetry and differential pulse voltammetry.  相似文献   

20.
The effects of lyotropic (swelling) anions (Cl(-), Br(-), NO(3)(-) and I(-)) on contractile properties of fast-twitch extensor digitorum longus (EDL) and slow-twitch soleus (SOL) muscles were investigated in vitro at 20 degrees C and 35 degrees C. Isolated muscles bathed in anionic Tyrode solution were stimulated directly and isometric single twitches and fused tetanic contractions were recorded. In a Cl(-)Tyrode solution a decrease of the bathing temperature led to a cold potentiation of the twitch tension (P(t)) in EDL muscles, however, to a cold depression in SOL muscles, in both muscles combined with a prolongation of contraction (CT) and half relaxation (HRT) times. The extent and order of the potentiating effect of lyotropic anions on the P(t), CT and HRT in EDL and SOL were quite similar and increased in the order: Cl(-)< Br(-)< NO(3)(-)< I(-). Since the lyotropic anions did not influence tetanic tensions, the twitch-tetanus ratio (TTR) was increased in NO(3)(-) and I(-)solutions. All effects of the anions were rapidly and completely reversed in both muscles when the test solution was replaced by the normal one. The temperature decrease caused no significant alteration in the potentiation capacity of the anions or in the kinetics of their action and reversibility.  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