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 52 毫秒
1.
Bernhardt PV  Schenk G  Wilson GJ 《Biochemistry》2004,43(32):10387-10392
Cyclic voltammetry of the non-heme diiron enzyme porcine purple acid phosphatase (uteroferrin, Uf) has been reported for the first time. Totally reversible one-electron oxidation responses (FeIII-FeII --> FeII-FeIII) are seen both in the absence and in the presence of weak competitive inhibitors phosphate and arsenate, and dissociation constants of these oxoanion complexes formed with uteroferrin in its oxidized state (Uf(o)) have been determined. The effect of pH on the redox potentials has been investigated in the range 3 < pH < 6.5, enabling acid dissociation constants for Uf(o) and its phosphate and arsenate complexes to be calculated.  相似文献   

2.
The oxidation-reduction potentials of ferredoxin-NADP+ reductase and flavodoxin from the cyanobacterium Anabaena PCC 7119 were determined by potentiometry. The potentials at pH 7 for the oxidized flavodoxin/flavodoxin semiquinone couple (E2) and the flavodoxin semiquinone/hydroquinone couple (E1) were -212 mV and -436 mV, respectively. E1 was independent of pH above about pH 7, but changed by approximately -60 mV/pH below about pH 6, suggesting that the fully reduced protein has a redox-linked pKa at about 6.1, similar to those of certain other flavodoxins. E2 varied by -50 mV/pH in the range pH 5-8. The redox potential for the two-electron reduction of ferredoxin-NADP+ reductase was -344 mV at pH 7 (delta Em = -30 mV/pH). In the 1:1 electrostatic complex of the two proteins titrated at pH 7, E2 was shifted by +8 mV and E1 was shifted by -25 mV; the shift in potential for the reductase was +4 mV. The potentials again shifted following treatment of the electrostatic complex with a carbodiimide, to covalently link the two proteins. By comparison with the separate proteins at pH 7, E2 for flavodoxin shifted by -21 mV and E1 shifted by +20 mV; the reductase potential shifted by +2 mV. The potentials of the proteins in the electrostatic and covalent complexes showed similar pH dependencies to those of the individual proteins. Qualitatively similar changes occurred when ferredoxin-NADP+ reductase from Anabaena variabilis was complexed with flavodoxin from Azotobacter vinelandii. The shifts in redox potential for the complexes were used with previously determined values for the dissociation constant (Kd) of the electrostatic complex of the two oxidised proteins, in order to estimate Kd values for the interaction of the different redox forms of the proteins. The calculations showed that the electrostatic complexes, formed when the proteins differ in their redox states, are stronger than those formed when both proteins are fully oxidized or fully reduced.  相似文献   

3.
Tan JH  Saint DA 《Life sciences》2000,67(22):2759-2766
Brief extracellular application of millimolar concentrations of lidocaine affected sodium currents recorded in isolated rat ventricular myocytes in two ways: 1) a reduction of the maximum current consistent with a channel blocking action, and 2) a negative shift in the voltage dependence of inactivation consistent with an interaction with the inactivated state of the channel. Both effects occurred very rapidly (< 1 s). Decreasing extracellular pH to 6.4 increased the potency for channel block (EC50 1.8 +/- 0.2 mM at pH 7.4 and 0.8 +/- 0.1 mM at pH 6.5) and decreased the potency to shift inactivation (V(1/2) shift -42 mV by 1 mM lidocaine at pH 7.4 and -12.6 mV at pH 6.5). Channel block was slightly less at +90 mV compared to -40 mV at either pH (not statistically significant). The increase in potency for block at decreased extracellular pH, while intracellular pH is buffered, and the lack of voltage dependence of block, suggest that the charged form of lidocaine can block the channel by interacting with a site near the extracellular mouth, although alternative explanations are discussed.  相似文献   

4.
The EPR spectra of the high-salt form of reduced bovine spleen purple acid phosphatase (BSPAPr) and its complexes with inhibitory tetrahedral oxyanions, AMP, and fluorine have been examined in the 4-30 K temperature range. The EPR spectrum of the high-salt form of BAPAPr is identical to that previously reported for the low-salt form (Averill et al. (1987) J. Am. Chem. Soc. 109, 3760-3767), indicating that the substantial differences in conformation of the two forms result in undetectable alterations in the electronic structure of the binuclear iron center. Phosphate, AMP, and arsenate all result in broadened, highly anisotropic EPR spectra with decreased values of the antiferromagnetic coupling constant, -2J, while molybdate and tungstate produce a sharp axial or slightly rhombic spectrum, respectively, and fluoride produces an anomalous spectrum with an inverted g-tensor. These results are consistent with binding of the two classes of oxyanions (and AMP) to distinct sites at or near the binuclear iron center, while fluoride binds in yet a third mode. EPR spectra of the BSPAPr complex with molybdate show altered relaxation behavior in the presence of phosphate, consistent with a 50% decrease in the magnitude of -2J, suggesting that phosphate binds to the molybdate complex to produce a ternary complex analogous to that proposed for molybdate inhibition on the basis of kinetics studies.  相似文献   

5.
Use of a revised purification procedure that maintains the enzyme in a high-salt environment has resulted in the isolation of a new form of the bovine spleen purple acid phosphatase. This enzyme cannot be distinguished from that previously described [Davis, J. C., Lin, S. S., & Averill, B. A. (1981) Biochemistry 20, 4062] by electrophoresis, isoelectric focusing, Western blot analysis, or N-terminal amino acid sequence and exhibits identical catalytic properties and EPR spectra in the reduced (pink) form. It does, however, possess a much more highly ordered structure as shown by CD spectra and exhibits markedly different reactivity upon oxidation and different visible spectra upon binding of inhibitory anions or changing pH. The properties of the new high-salt-stabilized form of the enzyme have permitted an extensive examination of the visible absorption spectra of complexes of the oxidized and reduced enzyme with inhibitory anions. It is found that these anions may be grouped into three classes on the basis of their effect on the visible absorption maximum and their sensitivity to pH: phosphate, arsenate, and AMP; tungstate and molybdate; and fluoride. This grouping is reinforced by a detailed examination of the steady-state kinetics of the enzyme in the presence of these inhibitors, which reveals that the first class exhibits mixed-type inhibition due to the presence of competitive and noncompetitive binding sites, while the second class exhibits simple non-competitive inhibition. Fluoride exhibits complex inhibition behavior characterized by curved Lineweaver-Burk plots; this behavior cannot be attributed to the presence of inhibitory aluminum fluoride complexes. Taken together, the spectral and kinetics data are consistent with a picture in which tetrahedral oxyanions bind in a noncompetitive fashion by bridging the two iron atoms in the dinuclear center, with the smaller anions also being able to bind in a competitive manner at a single iron atom.  相似文献   

6.
The redox potential dependence of the light-induced absorption changes of bacteriochlorophyll in chromatophores and subchromatophore pigment-protein complexes from Rhodospirillum rubrum has been examined. The highest values of the absorption changes due to the bleaching of P-870 and the blue shift of P-800 in chromatophores and subchromatophore complexes are observed in the 360-410mV redox potential range. At potentials below 300 mV (pH 7.0), the 880 nm band of bacteriochlorophyll shifts to shorter wavelengths in subchromatophore complexes and to longer wavelengths in chromatophores. The data on redox titration show that the red and blue shifts of 880-nm bacteriochlorophyll band represent the action of a non-identified component (C340) which has an oxidation-reduction midpoint potential close to 340 mV (n=1) at pH 6.0--7.6. The Em of this component varies by 60 mV/pH unit between pH 7.6 and 9.2. The results suggest that the red shift is due to the transmembrane, and the blue shift to the local intramembrane electrical field. The generation of both the transmembrane and local electrical fields is apparently governed by redox transitions of the component C340.  相似文献   

7.
Bacterial nitric oxide reductase (NOR) catalyzes the two-electron reduction of nitric oxide to nitrous oxide. It is a highly diverged member of the superfamily of heme-copper oxidases. The main feature by which NOR is distinguished from the heme-copper oxidases is the elemental composition of the active site, a dinuclear center comprised of heme b(3) and non-heme iron (Fe(B)). The visible region electronic absorption spectrum of reduced NOR exhibits a maximum at 551 nm with a distinct shoulder at 560 nm; these are attributed to Fe(II) heme c (E(m) = 310 mV) and Fe(II) heme b (E(m) = 345 mV), respectively. The electronic absorption spectrum of oxidized NOR exhibits a characteristic shoulder around 595 nm that exhibits complex behavior in equilibrium redox titrations. The first phase of reduction is characterized by an apparent shift of the shoulder to 604 nm and a decrease in intensity. This is due to reduction of Fe(B) (E(m) = 320 mV), while the subsequent bleaching of the 604 nm band represents reduction of heme b(3) (E(m) = 60 mV). This separation of redox potentials (>200 mV) allows the enzyme to be poised in the three-electron reduced state for detailed spectroscopic examination of the Fe(III) heme b(3) center. The low midpoint potential of heme b(3) represents a thermodynamic barrier to the complete (two-electron) reduction of the dinuclear center. This may avoid formation of a stable Fe(II) heme b(3)-NO species during turnover, which may be an inhibited state of the enzyme. It would also appear that the evolution of significant oxygen reducing activity by heme-copper oxidases was not simply a matter of the substitution of copper for non-heme iron in the dinuclear center. Changes in the protein environment that modulate the midpoint redox potential of heme b(3) to facilitate both complete reduction of the dinuclear center (a prerequisite for oxygen binding) and rapid heme-heme electron transfer were also necessary.  相似文献   

8.
The redox potential dependence of the light-induced absorption changes of bacteriochlorophyll in chromatophores and subchromatophore pigment-protein complexes from Rhodospirillum rubrum has been examined. The highest values of the absorption changes due to the bleaching of P-870 and the blue shift of P-800 in chromatophores and subchromatophore complexes are observed in the 360–410 mV redox potential range. At potentials below 300 mV (pH 7.0), the 880 nm band of bacteriochlorophyll shifts to shorter wavelengths in subchromatophore complexes and to longer wavelengths in chromatophores.

The data on redox titration show that the red and blue shifts of 880-nm bacteriochlorophyll band represent the action of a non-identified component (C340) which has an oxidation-reduction midpoint potential close to 340 mV (n = 1) at pH 6.0–7.6. The Em of this component varies by 60 mV/pH unit between pH 7.6 and 9.2.

The results suggest that the red shift is due to the transmembrane, and the blue shift to the local intramembrane electrical field. The generation of both the transmembrane and local electrical fields is apparently governed by redox transitions of the component C340.  相似文献   


9.
L D Byers  H S She  A Alayoff 《Biochemistry》1979,18(12):2471-2480
The glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase catalyzes the oxidative phosphorylation of D-glyceraldehyde 3-phosphate. A variety of phosphonates have been shown to substitute for phosphate in this reaction [Gardner, J. H., & Byers, L. D., (1977) J. Biol. Chem. 252, 5925--5927]. The dependence of the logarithm of the equilibrium constant for the reaction on the pKa2 value of the phosphonate is characterized by a Br?nsted coefficient, betaeq, of approximately 1. This represents the sensitivity of the transfer of the phosphoglyceroyl group between the active-site sulfhydryl residue (in the acyl-enzyme intermediate) and the acyl acceptor on the basicity of the acyl acceptor. Molybdate (MoO42-) can also serve as an acyl acceptor in the glyceraldehyde-3-phosphate dehydrogenase catalyzed reaction. The second-order rate constant for the reaction with molybdate is only approximately 12 times lower than the reaction with phosphate even though the pKa2 of molybdate is 3.1 units lower than the pKa2 of phosphate. The immediate product of the molybdate reaction is the acyl molybdate, 1-molybdo-3-phosphoglycerate. The acyl molybdate, like the acyl arsenate (the immediate product of the reaction when arsenate is the acyl acceptor), is kinetically unstable. At pH 7.3 (25 degrees C), the half-life for hydrolysis of the acyl molybdate, or the acyl arsenate, is less than 2.5 s. Thus, hydrolysis of 1-molybdo- and 1-arseno-3-phosphoglycerate is at least 2000 times faster than hydrolysis of 1,3-diphosphoglycerate under the same conditions. Glyceraldehyde-3-phosphate dehydrogenase has a fairly broad specificity for acyl acceptors. Most tetrahedral oxy anions tested are substrates for the enzyme (except SO4(2-) and SeO4(2-)). Tetrahedral monoanions such as ReO4- and GeO(OH)3- are not substrates but do bind to the enzyme. These results suggest the requirement of at least one anionic site on the acyl acceptor required for binding and another anionic group on the acyl receptor required for nucleophilic attack on the acyl enzyme.  相似文献   

10.
Zu Y  Couture MM  Kolling DR  Crofts AR  Eltis LD  Fee JA  Hirst J 《Biochemistry》2003,42(42):12400-12408
Rieske [2Fe-2S] clusters can be classified into two groups, depending on their reduction potentials. Typical high-potential Rieske proteins have pH-dependent reduction potentials between +350 and +150 mV at pH 7, and low-potential Rieske proteins have pH-independent potentials of around -150 mV at pH 7. The pH dependence of the former group is attributed to coupled deprotonation of the two histidine ligands. Protein-film voltammetry has been used to compare three Rieske proteins: the high-potential Rieske proteins from Rhodobacter sphaeroides (RsRp) and Thermus thermophilus (TtRp) and the low-potential Rieske ferredoxin from Burkholderia sp. strain LB400 (BphF). RsRp and TtRp differ because there is a cluster to serine hydrogen bond in RsRp, which raises its potential by 140 mV. BphF lacks five hydrogen bonds to the cluster and an adjacent disulfide bond. Voltammetry measurements between pH 3 and 14 reveal that all the proteins, including BphF, have pH-dependent reduction potentials with remarkably similar overall profiles. Relative to RsRp and TtRp, the potential versus pH curve of BphF is shifted to lower potential and higher pH, and the pK(a) values of the histidine ligands of the oxidized and reduced cluster are closer together. Therefore, in addition to simple electrostatic effects on E and pK(a), the reduction potentials of Rieske clusters are determined by the degree of coupling between cluster oxidation state and histidine protonation state. Implications for the mechanism of quinol oxidation at the Q(O) site of the cytochrome bc(1) and b(6)f complexes are discussed.  相似文献   

11.
The sensitivity to Ca2+ of the Ca2+-dependent K+ channel can be increased by the artificial electron donor system ascorbate + phenazine-methosulphate in a variety of animal cells. In the human erythrocyte the shift from the 'low' to the 'high-affinity' state seems to depend on the reduction of a membrane component accepting 2 electrons and with an standard redox potential (pH 7.5) of about 47 mV. The relevance of this redox modulation under physiological circumstances is unknown at the moment.  相似文献   

12.
Properties of the fully developed phosphate transport system in the fertilized egg of the sea urchin, Strongylocentrotus purpuratus, were investigated. The rates of phosphate transport at concentrations of external phosphate of 1 to 44 μM, both in the absence and in the presence of 100 μM arsenate, exhibit typical saturation kinetics. At sea water concentrations of 2 μM phosphate, the rate of uptake is about 2 × 10?9 μm/egg/minute at 15°C. Arsenate is a competitive inhibitor of phosphate transport, fully and immediately reversible in its effects, yielding Ki values ranging from 10.5 to 14.1 × 10?6 M in comparison to the corresponding apparent KM (Michaelis-Menten) constants for phosphate of 5.6 to 7.5 × 10?6 M (pH 8.0, 15°C). The rate of arsenate uptake in a phosphate deficient medium amounts to 2.8 to 2.9 × 10?10 μm arsenate/egg/minute at an arsenate concentration of 2.9 to 10.2 μM arsenate (HAsO4??), which is 9.5 and 5.6% of the rate of phosphate uptake at corresponding phosphate concentrations. Arsenate has essentially the same developmental effects at initial concentrations of 5–10 μM and 100 μM arsenate, namely no observable effects for exposure periods of 7.5 hours, although longer periods result in blockage of development at the early blastula stage. Outward flux of phosphate ions cannot be demonstrated by washing prelabelled eggs with sea water containing low or high concentrations of phosphate, even when phosphorylation has been blocked by exposing the eggs to a metabolic inhibitor. Phosphate uptake rates measured in the pH range from 5.0 to 10.0 reveal a sharp optimum at pH 8.8–8.9. Reference to the apparent pK' values of the phosphoric acid system indicate that the entering species is the HPO4?? ion. The effects on rates of phosphate uptake of exposure to sea water at pH values between 7 and 10 for 30 minute periods are fully reversible, but at lower pH values, reversal is delayed, and is only partial. Sodium molybdate (0.01 M), sodium pyrophosphate (1.5 × 10?4 M), and adenosine triphosphate (1–5 × 10?4 M) for exposure periods ranging from 40 to 180 minutes did not significantly affect phosphate uptake. Omission of Ca++ ion from artificial sea water is without effect on phosphate uptake but the absence of both Ca++ and Mg++ results in profound and irreversible depression of both phosphate uptake and development. The data of this and the following paper are consistent with the conclusion that the transport of phosphate involves a surface located carrier. The apparent secondary and tertiary ionization constants of phosphoric acid in sea water (ionic strength = 0.6885) were measured, resulting in a value for pK′2 = 6.14 and for pK′3 = 10.99, at 15°C and phosphate at infinite dilution.  相似文献   

13.
The diiron active sites of the purple acid phosphatases from porcine uterus (also called uteroferrin, Uf) and bovine spleen (BSPAP) and their complexes with tungstate are compared by 1H NMR and NOE techniques. The paramagnetically shifted features of the 1H NMR spectrum of reduced BSPAP are similar to those of reduced Uf, while the spectra of the tungstate complexes are almost identical. These observations suggest that the two active sites are quite similar, in agreement with the greater than 90% sequence homology found in the two enzymes. Nuclear Overhauser effect (NOE) experiments on the His N-H resonances show that the Fe(III)-His residue is N epsilon-coordinated, while the Fe(II)-His is H delta-coordinated in both enzymes. On the basis of the above NMR and NOE results, our previously proposed model for the dinuclear iron active site of Uf [Scarrow, R. C., Pyrz, J. W., & Que, L., Jr. (1990) J. Am. Chem. Soc. 112, 657-665] is corroborated, refined, and found to represent the diiron center of BSPAP as well.  相似文献   

14.
Harold, F. M. (National Jewish Hospital, Denver, Colo.), and J. R. Baarda. Interaction of arsenate with phosphate-transport systems in wild-type and mutant Streptococcus faecalis. J. Bacteriol. 91:2257-2262. 1966.-Arsenate competitively inhibits the growth of Streptococcus faecalis, primarily by competition with phosphate for a common transport system. Arsenate is itself accumulated by the cells; the uptake requires metabolic energy, and the intracellular arsenate level may reach 0.01 m. Cells loaded with arsenate have lost the capacity to take up radioactive glutamate, rubidium, phosphate, or arsenate itself, apparently by the uncoupling of adenosine triphosphate generation. The pH dependence of arsenate uptake is complex. At low concentrations of extracellular arsenate, uptake by the wild-type strain 9790 exhibits a single maximum about pH 8; mutant PT-1, previously shown to be defective in phosphate uptake, takes up essentially no arsenate. At high concentrations of arsenate, uptake by the wild type is bimodal with maxima at pH 5.5 and 9; the uptake curve for mutant PT-1 corresponds to the shoulder in the curve for the wild type. The apparent dissociation constant for arsenate uptake by the wild type is approximately 10(-5)m from pH 5 to 9, whereas that for mutant PT-1 is about 5 x 10(-5) M at pH 5 and rises rapidly with increasing pH. The results confirm the earlier conclusion that the lesion in mutant PT-1 resides in the transport of phosphate and arsenate. It is proposed that the wild type has two distinct transport systems, whereas the mutant has lost the one with alkaline pH optimum.  相似文献   

15.
The acid phosphate activity (APA) associated with the isolated brush border membrane of the tapeworm, Hymenolepis diminuta, hydrolyzed p-nitrophenyl phosphate (PNPP), pyrophosphate (PPi), and beta-glycerophosphate (beta GP). Inhibition of PNPP hydrolysis at pH 4.0 was inhibited in a competitive manner by the following compounds (listed in order of decreasing affinity with their apparent inhibitor constants (Ki')): molybdate (0.031 mM); PPi (0.147 mM); NaF (0.150 mM); o-carboxyphenyl phosphate (0.261 mM); inorganic phosphate (0.770)); arsenate (3.45 mM); tartrate (22.1 mM); and beta GP (29.8 mM). Cu2+, formaldehyde, and arsenite at 10:1, 80:1, and 200:1 inhibitor to substrate ratios did not inhibit APA. The maximal rate of hydrolysis (Vmax) of each substrate was greater at pH 4.0 than 5.0. The apparent Michaelis constant (Km') for PNPP increased from 0.233 to 0.351 mM when the pH was raised from 4.0 to 5.0. The Km' for PPi decreased from 0.101 to 0.046 mM, while the Km' for beta GP changed from 2.04 to 2.22 mM under similar circumstances. APA and alkaline phosphatase activity increased as a function of temperature up to 45 degrees C.  相似文献   

16.
1. Acid phosphatase from the yeast Rhodotorula rubra was purified 44-fold. The purification procedure involved mechanical disruption of cells, precipitation with ethanol, chromatography on DEAE- and CM-cellulose. 2. The purified enzyme is homogeneous in polyacrylamide gels at pH 4.5, 9.5 and 8.4. Carbohydrate content accounts for 57% of the total weight. The optimum pH is at 4.0-4.6, and the enzyme is stable over pH range from 2.6 to 6.0. Full activity was retained on 60-min incubation at 50 degrees C, but it was reduced by half on 60-min incubation at 65 degrees C. 3. Specificity of the enzyme is fairly broad; monoesters of carbohydrates, and nucleosides and inorganic pyrophosphate can serve as substrates. Km was found to be 1 X 10(-4) M for p-nitrophenyl phosphate as a substrate. The enzyme is inhibited by molybdate, phosphate, arsenate and fluoride ions.  相似文献   

17.
The reaction between cytochrome c oxidase and ferrocytochrome c has been investigated by the stopped-flow method. It has been found that only one electron acceptor, a heme group, in the oxidase is rapidly reduced by cytochrome c. The presence of N3- does not affect the reduction of the acceptor, which supports the hypothesis that this is identical with cytochrome a. The results are consistent with the existence of a simple equilibrium between cytochrome a and cytochrome c: c-2 + a-3+ in equilibrium c-3+ + a-2+ with an equilibrium constant corresponding to an oxidation-reduction potential of cytochrome a 30 mV higher than that for cytochrome c at pH 7.4. The oxidation-reduction potential of the a-3+ /a-2+ couple, 285 mV (based on a potential of 255 mV for cytochrome c), and the optical properties of the reduced form indicate that it is identical with neither of the reduced hemes seen in potentiometric titrations. The oxidase species resulting from the rapid reduction of cytochrome a by cytochrome c is proposed to represent a metastable intermediate state which, under anaerobic conditions, eventually is transformed into a more stable state characterized by a reduced high-potential heme.  相似文献   

18.
Interference of arsenate and vanadate with phosphate uptakein Lemna gibba L. was studied by measuring voltage changes and(32P)phosphate uptake. Arsenate proved to be competitive withthe high- and low-affinity phosphate uptake system. It inducedtransient membrane potential changes of up to 120 mV which weresimilar to those induced by phosphate and indicated a cotransportmechanism with at least 2H+/H2As. The amplitude of the transient arsenate-induced membrane depolarization wasstrongly influenced by phosphate starvation. A permanent membranedepolarization to the diffusion potential was achieved within2 to 6 h in P-starved plants. Thus, arsenate is indeed a stronglycompetitive physiological analogue of phosphate in higher plants. Vanadate was easily transported into L. gibba as concluded fromtransient Em changes of up to 110 mV. Vanadate interfered onlyslightly and non-specifically with the two phosphate transportmechanisms. Like phosphate, vanadate uptake seems to be an H+-cotransportmechanism, both with similar optima at pH 6.0. Unlike phosphateuptake, vanadate-linked membrane depolarization was not affectedby high intracellular phosphate concentrations. P-starvationdid not enhance the weak long-term effect on Em. Hence, vanadate,in contrast to arsenate, is not regarded as a physiologicalphosphate analogue. The distinct and rapid vanadate-inducedand permanent membrane depolarization of Avena sativa, Triticumaestivum and Glycine max leaves was not seen in Lemna nor inleaves of Gossypium hirsutum and Nicotiana tabacum. Plasmalemma-enrichedpreparations of L. gibba revealed, however, a high vanadate-sensitiveATPase activity (87%). As a possible explanation for these differencesit is suggested that the latter plant species have cytosolicvanadate-detoxifying properties, i.e. they can reduce vanadateto vanadyl ions, in contrast to the former group of plant species. Key words: Arsenate, vanadate, H+/solute cotransport, membrane potential, phosphate competition  相似文献   

19.
The steady-state gating of individual batrachotoxin-modified sodium channels in neutral phospholipid bilayers exhibits spontaneous, reversible changes in channel activation, such that the midpoint potential (Va) for the gating curves may change, by 30 mV or more, with or without a change in the apparent gating valence (za). Consequently, estimates for Va and, in particular, za from ensemble-averaged gating curves differ from the average values for Va and za from single-channel gating curves. In addition to these spontaneous variations, the average Va shifts systematically as a function of [NaCl] (being -109, -88, and -75 mV at 0.1, 0.5, and 1.0 M NaCl), with no systematic variation in the average za (approximately 3.7). The [NaCl]-dependent shifts in Va were interpreted in terms of screening of fixed charges near the channels' gating machinery. Estimates for the extracellular and intracellular apparent charge densities (sigma e = -0.7 and sigma i = -0.08 e/nm2) were obtained from experiments in symmetrical and asymmetrical NaCl solutions using the Gouy-Chapman theory. In 0.1 M NaCl the extracellular and intracellular surface potentials are estimated to be -94 and -17 mV, respectively. The intrinsic midpoint potential, corrected for the surface potentials, is thus about -30 mV, and the standard free energy of activation is approximately -12 kJ/mol. In symmetrical 0.1 M NaCl, addition of 0.005 M Ba2+ to the extracellular solution produced a 17-mV depolarizing shift in Va and a slight reduction in za. The shift is consistent with predictions using the Gouy-Chapman theory and the above estimate for sigma e. Subsequent addition of 0.005 M Ba2+ to the intracellular solution produced a approximately 5-mV hyperpolarizing shift in the ensemble-averaged gating curve and reduced za by approximately 1. This Ba(2+)-induced shift is threefold larger than predicted, which together with the reduction in za implies that Ba2+ may bind at the intracellular channel surface.  相似文献   

20.
Hoke KR  Cobb N  Armstrong FA  Hille R 《Biochemistry》2004,43(6):1667-1674
Arsenite oxidase from Alcaligenes faecalis, an unusual molybdoenzyme that does not exhibit a Mo(V) EPR signal during oxidative-reductive titrations, has been investigated by protein film voltammetry. A film of the enzyme on a pyrolytic graphite edge electrode produces a sharp two-electron signal associated with reversible reduction of the oxidized Mo(VI) molybdenum center to Mo(IV). That reduction or oxidation of the active site occurs without accumulation of Mo(V) is consistent with the failure to observe a Mo(V) EPR signal for the enzyme under a variety of conditions and is indicative of an obligate two-electron center. The reduction potential for the molybdenum center, 292 mV (vs SHE) at pH 5.9 and 0 degrees C, exhibits a linear pH dependence for pH 5-10, consistent with a two-electron reduction strongly coupled to the uptake of two protons without a pK in this range. This suggests that the oxidized enzyme is best characterized as having an L(2)MoO(2) rather than L(2)MoO(OH) center in the oxidized state and that arsenite oxidase uses a "spectator oxo" effect to facilitate the oxo transfer reaction. The onset of the catalytic wave observed in the presence of substrate correlates well with the Mo(VI/IV) potential, consistent with catalytic electron transport that is limited only by turnover at the active site. The one-electron peaks for the iron-sulfur centers are difficult to observe by protein film voltammetry, but spectrophotometric titrations have been carried out to measure their reduction potentials: at pH 6.0 and 20 degrees C, that of the [3Fe-4S] center is approximately 260 mV and that of the Rieske center is approximately 130 mV.  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