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 41 毫秒
1.
Haim Garty  S.Roy Caplan 《BBA》1977,459(3):532-545
The uptake of rubidium in intact Halobacterium halobium cells was followed, and found to be light-dependent. The exchange process is slow, the steady-state rate of 86Rb+/Rb+ exchange being given by k = 6.3 · 10?4 min?1. Starved cells exhibited a faster rate than unstarved cells. The influx of 86Rb+ was almost completely blocked in the presence of proton conductors (CCCP, FCCP, and SF 6847), and was sensitive to the presence of the permeant cation TPMP+. Valinomycin very slightly increased the rate of uptake, while 1 · 10?6 M nigericin showed significant inhibition. On the other hand, release of 86Rb+ was not light-dependent, although still affected by uncouplers, TPMP+, and nigericin. These experimental observations may be explained in terms of a passive flux driven by an electrical potential difference, and influenced by positive isotope interaction within the membrane. In carefully matched influx-efflux studies, the extent of the positive isotope interaction was measured. Using the formal treatment of Kedem and Essig, the ratio (exchange resistance)/(resistance to net flow) for 86Rb+ was found to be 1.7.  相似文献   

2.
The kinetics of interfacial proton transfer reaction is an important factor in proton transport across membranes. The following experimental system was designed in order to measure this kinetics. Sonicated liposomes having the protonophore SF6847 was suspended in Tris buffer. Application of a temperature jump (in ∼ 3 μs) caused a drop in the aqueous phase pH which was subsequently sensed by the membrane-bound SF6847. The kinetics of this interfacial proton transfer reaction was monitored on μs timescales. The estimated bimolecular rate constant of 2×1011 M−1 s#x2212;1 for this process show that there is no kinetic barrier for the transfer of protons from the aqueous phase to the membrane-water interface.  相似文献   

3.
Evert P. Bakker  S.Roy Caplan 《BBA》1978,503(2):362-379
The method of Warren et al. (1974, Proc. Natl. Acad. Sci. U.S. 71, 622–626) was employed to substitute the polar lipids of the purple membrane of Halobacterium halobium by different phosphatidylcholine species. Substitution at pH 6.5 yields proteolipid complexes in the form of bent open sheets which have a protein to lipid phosphorus ratio similar to the natural membrane, i.e. about 1 : 10 (mol/mol). The extent of substitution increases with the length of the fatty acid chain of the phosphatidylcholine used.The spectral properties of bacteriorhodopsin are only slightly affected by substitution of 95% of the lipid, except that the photocycle is slowed down appreciably. Due to this slow rate the M4 12 intermediate of the cycle accumulates in the light. Associated with this accumulation is a net light-induced proton release, which proved insensitive to uncoupler. A comparison between the net proton release and the amount of M4 12 accumulated, studied as a function of pH, shows that no fixed stoichiometry exists between the two processes.Phospholipid substitution by egg phosphatidylcholine at pH 7.5 or by egg phosphatidylethanolamine leads to preparations of purple membrane with 15 or 25 mol of phospholipid per mol of bacteriorhodopsin, respectively. These preparations seem to consist of closed membrane structures. They take up protons in the light in an uncoupler-sensitive way.  相似文献   

4.
The rate of proton transfer between the octanol -OH group and water dissolved in octanol after partition equilibrium was determined by 1H-NMR spectrometry. The rate was found to depend on the pH of the aqueous phase, being minimal at about pH 11. The uncoupler of oxidative phosphorylation 2,4-dinitrophenol at about 10?3 M accelerated proton transfer several-fold. Its effect was shown to depend on the concentration of the neutral form of 2,4-dinitrophenol in the octanol phase, irrespective of the pH of the aqueous phase. This effect is suggested to be based on the catalytic action of the phenolic -OH group in 2,4-dinitrophenol. The importance of this effect in the uncoupling action of 2,4-dinitrophenol is discussed.  相似文献   

5.
Membrane vesicles from a red mutant of Halobacteriumhalobium R1 accumulate protons when illuminated causing the pH of the suspension to rise. Sodium is extruded from the vesicles and a membrane potential is formed. This potential and the proton uptake are abolished by valinomycin if K+ is present. In contrast, Na+-efflux is uninhibited by valinomycin even though no membrane potential is detectable and H+ influx does not occur. Bis (hexafluoracetonyl)acetone (1799) stimulates proton uptake but does not abolish membrane potential. We propose that a light-dependent sodium pump is present. Passive proton uptake occurs in response to the electrical gradient created by this light-driven Na+ pump in contrast to the active proton, and passive Na+ flux that occurs in response to the light-driven proton pump described in vesicles of the parent strain of H.halobium R1.  相似文献   

6.
The effects of N,N′-dicyclohexylcarbodiimide (DCCD), triphenyltin chloride (TPT), and 3,5-di-tert-butyl-4-hydroxybenzylidenemalonomtrile (SP6847) were tested on the light-dependent activities of Halobacterium halobium R1mR which contains a new retinal protein pigment designated as halorhodopsin but no bacteriorhodospin. DCCD inhibited ATP synthesis either in the light- or in the dark-aerobic conditions without affecting the light-induced proton uptake (ΔH+). Although DCCD lowered the membrane potential under dark-anaerobic conditions, the potential increased in the light as high as the control (the light-dependent membrane potential increment Δψ became apparently larger in the presence of DCCD). TPT had negligible effect on ATP synthesis both in the dark or in the light but inhibited markedly ΔH+ and partly Δψ. After R1mR was treated with DCCD, TPT abolished ΔH+ almost completely but Δψ only partly. The remaining Δψ was collapsed by SF6847 with a concomitant proton incorporation (pH increase). These results led to the following postulations: (i) In R1mR, ATP is synthesized by a H+-ATPase coupled either to respiration and/or light energization by halorhodopsin; (ii) the majority of protons are incorporated in the light by a mechanism which differs from H+-ATPase but is driven by the Δψ generated by halorhodopsin; (iii) TPT acts in this system as a chloride/hydroxide exchanger; (iv) the uncoupler SF6847 carries protons into cells in response to Δψ.  相似文献   

7.
1. Modification of the Class II sulphydryl groups on the (Na+ + K+)-ATPase from rectal glands of Squalus acanthias with N-ethylmaleimide has been used to detect conformational changes in the protein. The rates of inactivation of the enzyme and the incorporation of N-ethylmaleimide depend on the ligands present in the incubation medium. With 150 mM K+ the rate of inactivation is largest (k1 = 1.73 mM?1 · min?1) and four SH groups per α-subunit are modified. The rate of inactivation in the presence of 150 mM Na+ is smaller (k1 = 1.08 mM?1 · min-1) but the incorporation of N-ethylmaleimide is the same as with K+. 2. ATP in micromolar concentrations protects the Class II groups in the presence of Na+ (k1 = 0.08 mM?1 · min?1 at saturating ATP) and the incorporation id drastically reduced. ATP in millimolar concentrations protects the Class II groups partially in the presence of K+ (k1 = 1.08 mM?1 · min?1) and three SH groups are labelled per α subunit. 3. The K+ -dependent phosphatase is inhibited in parallel to the (Na+ + K+)-ATPase under all conditions, and the ligand-dependent incorporation of N-ethylmaleimide was on the α-subunit only. 4. It is shown that the difference between the Na+ and K+ conformations sensed with N-ethylmaleimide depends on the pH of the incubation medium. At pH 6 there is a very small difference between the rates of inactivation in the presence of Na+ and K+, but at higher pH the difference increases. It is also shown that the rate of inactivation has a minimum at pH 6.9, which suggests that the conformation of the enzyme changes with pH. 5. Modification of the Class III groups with N-ethylmaleimide-whereby the enzyme activity is reduced from about 16% to zero-shows that these groups are also sensitive to conformational changes. As with the Class II groups, ATP in micromolar concentrations protects in the presence of Na+ relative to Na+ or K+ alone. ATP in millimolar concentrations with K+ present increases the rate of inactivation relative to K+ alone, in contrast to the effect on the Class II groups. 6. Modification of the Class II groups with a maleimide spin label shows a difference between Class II groups labelled in the presence of Na+ (or K+) and Class II groups labelled in the presence of K + ATP, in agreement with the difference in incorporation of N-ethylmaleimide. The spectra suggest that the SH group protected by ATP in the presence of K+ is buried in the protein. 7. The results suggest that at least four different conformations of the (Na+ + K+)-ATPase can be sensed with N-ethylmaleimide: (i) a Na+ form of the enzyme with ATP bound to a high-affinity site (E1-Na-ATP); (ii) a Na+ form without ATP bound (E1-Na); (iii) a K+ form without ATP bound (E2-K); and (iv) an enzyme form with ATP bound to a low-affinity site in the presence of K+, probably and E1-K-ATP form.  相似文献   

8.
A highly purified membrane fraction was derived from hog gastric mucosa by a combination of differential and density gradient centrifugation and free flow electrophoresis. This final fraction was 35-fold enriched with respect to cation activated ouabain-insensitive ATPase. Antibody against this fraction was shown to be bound to the luminal surface of the gastric glands. The addition of ATP to this fraction or the density gradient fraction resulted in H+ uptake into an osmotically sensitive space. The apparent Km for ATP was 1.7 · 10?4 M in the absence of a K+ gradient similar to that found for ATPase activity. The reaction is specific for ATP and requires cation in the sequence K+ > Rb+ > Cs+ > Na+ > Li+ and is inhibited by ATPase inhibitors such as N,N′-dicylclohexylcarbodiimide. Maximal H+ uptake occurs with an outward K+ gradient but the minimal apparent KA is found in the absence of a K+ gradient. The pH optimum for H+ uptake is between 5.8 and 6.2 which corresponds to the pH range for phosphorylation of the enzyme, but is considerably less than the pH maximum of the K+ dependent dephosphorylation. In the presence of an inward K? gradient, protonophores such as tetrachlorsalicylanilide only partially abolish the H+ gradient but valinomycin dissipates 75% of the gradient, and nigericin abolishes the gradient. The vesicles therefore have a low K+ conductance but a measurable H+ conductance, hence a K+ gradient can produce an H+ gradient in the presence of valinomycin. The uptake and spontaneous leak of H+ are temperature sensitive skin with a similar transition temperature. Ultraviolet irradiation inactivates ATPase and proton transport at the same rate, approximately at twice the rate of p-nitrophenylphosphatase inactivation. It is concluded that H+ uptake by these vesicles is probably due to a dimeric (H+ + K+)-ATPase and is probably non-electrogenic.  相似文献   

9.
Uptake of d-glucosamine by rat brain synaptosomes occurs via a saturable transport process (Km 2.1 mM, V 3.0 nmol/mg per min) which was clearly distinguishable from simple diffusion. This transport process is highly sensitive to cytochalasin (Ki = 7 · 10?5 mM. d-Glucose competitively inhibits d-glucosamine uptake with a Ki value of 8 · 10?1 mM.  相似文献   

10.
Pyranine is shown to be a convenient and sensitive probe for reporting pH values, pHi, at the interior of anionic and at the outer surface of cationic liposomes. It is well shielded from the phospholipid headgroups by water molecules in the interior of anionic liposomes, but it is bound to the surface of cationic liposomes. Hydrogen ion concentrations outside the liposomes, ‘bulk pH values’, pHo, were measured by a combination electrode. While pHi = pHo for neutral, pHi < pHo for anionic and pHi > pHo for cationic liposomes prepared in 5.0 · 10?3 M phosphate buffers. pKa values for the ionization of pyranine were 7.22 ± 0.04 and 6.00 ± 0.05 in water and at the external surface of cationic liposomes. The surface potential for cationic liposomes containing dipalmitoyl-d-α-phosphatidylcholine, cholesterol and octadecylamine in the molar ratio of 1.00 : 0.634 : 1.01, were calculated to be +72.2 mV. Proton permeabilities were measured for single and multicompartment anionic liposomes. Transfer of anionic liposomes prepared at a given pH to a solution of different pH resulted in a pH gradient if sodium phosphate or borate were used as buffers. In the presence of sodium acetate proton equilibration is promptly established.  相似文献   

11.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

12.
Neeraj Agarwal  Vijay K. Kalra 《BBA》1983,723(2):150-159
Interaction of N,N′-dicyclohexylcarbodiimide (DCCD) with ATPase of Mycobacterium phlei membranes results in inactivation of ATPase activity. The rate of inactivation of ATPase was pseudo-first order for the initial 30–65% inactivation over a concentration range of 5–50 μM DCCD. The second-order rate constant of the DCCD-ATPase interaction was k = 8.5·105 M?1·min?1. The correlation between the initial binding of [14C]DCCD and 100% inactivation of ATPase activity shows 1.57 nmol DCCD bound per mg membrane protein. The proteolipid subunit of the F0F1-ATPase complex in membranes of M. phlei with which DCCD covalently reacts to inhibit ATPase was isolated by labeling with [14C]DCCD. The proteolipid was purified from the membrane in free and DCCD-modified form by extraction with chloroform/methanol and subsequent chromatography on Sephadex LH-20. The polypeptide was homogeneous on SDS-acrylamide gel electrophoresis and has an apparent molecular weight of 8000. The purified proteolipid contains phosphatidylinositol (67%), phosphatidylethanolamine (18%) and cardiolipin (8%). Amino acid analysis indicates that glycine, alanine and leucine were present in elevated amounts, resulting in a polarity of 27%. Cysteine and tryptophan were lacking. Butanol-extracted proteolipid mediated the translocation of protons across the bilayer, in K+-loaded reconstituted liposomes, in response to a membrane potential difference induced by valinomycin. The proton translocation was inhibited by DCCD, as measured by the quenching of fluorescence of 9-aminoacridine. Studies show that vanadate inhibits the proton gradient driven by ATP hydrolysis in membrane vesicles of M. phlei by interacting with the proteolipid subunit sector of the F0F1-ATPase complex.  相似文献   

13.
The Caulobacter crescentus (NA1000) xynB5 gene (CCNA_03149) encodes a predicted β-glucosidase-β-xylosidase enzyme that was amplified by polymerase chain reaction; the product was cloned into the blunt ends of the pJet1.2 plasmid. Analysis of the protein sequence indicated the presence of conserved glycosyl hydrolase 3 (GH3), β-glucosidase-related glycosidase (BglX) and fibronectin type III-like domains. After verifying its identity by DNA sequencing, the xynB5 gene was linked to an amino-terminal His-tag using the pTrcHisA vector. A recombinant protein (95 kDa) was successfully overexpressed from the xynB5 gene in E. coli Top 10 and purified using pre-packed nickel-Sepharose columns. The purified protein (BglX-V-Ara) demonstrated multifunctional activities in the presence of different substrates for β-glucosidase (pNPG: p-nitrophenyl-β-D-glucoside) β-xylosidase (pNPX: p-nitrophenyl-β-D-xyloside) and α-arabinosidase (pNPA: p-nitrophenyl-α-L-arabinosidase). BglX-V-Ara presented an optimal pH of 6 for all substrates and optimal temperature of 50 °C for β-glucosidase and α-l-arabinosidase and 60 °C for β-xylosidase. BglX-V-Ara predominantly presented β-glucosidase activity, with the highest affinity for its substrate and catalytic efficiency (Km 0.24 ± 0.0005 mM, Vmax 0.041 ± 0.002 µmol min?1 mg?1 and Kcat/Km 0.27 mM?1 s?1), followed by β-xylosidase (Km 0.64 ± 0.032 mM, Vmax 0.055 ± 0.002 µmol min?1 mg?1 and Kcat/Km 0.14 mM?1s?1) and finally α-l-arabinosidase (Km 1.45 ± 0.05 mM, Vmax 0.091 ± 0.0004 µmol min?1 mg?1 and Kcat/Km 0.1 mM?1 s?1). To date, this is the first report to demonstrate the characterization of a GH3-BglX family member in C. crescentus that may have applications in biotechnological processes (i.e., the simultaneous saccharification process) because the multifunctional enzyme could play an important role in bacterial hemicellulose degradation.  相似文献   

14.
An extracellular acid phosphatase secreted into the medium during growth of Tetrahymena pryiformis strain W was purified about 900-fold by (NH4)2SO4 precipitation, gel filtration and ion exchange chromatography. The purified acid phosphatase was homogenous as judged by polycrylamide gel electrophoresis and was found to be a glycoprotein. Its carbohydrate content was about 10% of the total protein content. The native enzyme has a molecular weight of 120 000 as determined by gel filtration and 61 000 as determined by sodium dodecyl sulfate-polycrylamide gel electrophoresis. The acid phosphatase thus appears to consist of two subunits of equal size. The amino acid analysis revealed a relatively high content of asparic acid, glutamic acid and leucine. The purified acid phosphatase from Tetrahymena had a rather broad substrate specificity; it hydrolyzed organic phosphates, nucleotide phosphates and hexose phosphates, but had no diesterase activity. The Km values determined with p-nitrophenyl phosphate, adenosine 5′-phosphate and glucose 6-phosphate were 3.1·10?4 M, 3.9·10?4 M and 1.6·10?3 M, respectively. The optima pH for hydrolysis of three substrates were similar (pH 4.6). Hg2+ and Fe3+ at 5 mM were inhibitory for the purified acid phosphatase, and fluoride, L-(+)-tartaric acid and molybdate also inhibited its cavity at low concentrations. The enzyme was competitively inhibited by NaF (Ki=5.6·10?4 M) and by L-(+)-tartaric acid (Ki = 8.5·10?5 M), while it was inhibited noncompetitively by molybdate Ki = 5.0·10?6 M). The extracellular acid phosphatase purified from Tetrahymena was indistinguishable from the intracellular enzyme in optimum pH, Km, thermal stability and inhibition by NaF.  相似文献   

15.
Various physicochemical and biochemical properties of the most potent uncoupler of oxidative phosphorylation known to date 3,5-di-tert-butyl-4-hydroxybenzylidenemalononitrile (SF 6847), such as pH dependence of the uncoupling activity and binding to mitochondria, spectral properties in the presence of different types of liposomes, biopolymers and mitochondria, and effects on model membrane systems have been investigated. From the results, it is concluded that the uncoupler most likely is localized in the phospholipid part of the membrane.  相似文献   

16.
The interior of purified cholinergic Torpedo vesicles is acidic, pHin = 5.2 at external pH = 7.4. The internal pH changes linearily as a function of external pH yielding ΔpH = 2.0 and 2.5 at pHout = 6.3 and 9.1 respectively. The proton translocator carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) + the ionophore valinomycin dissipate the proton gradient across the vesicular membrane and concurrently induce acetylcholine release from vesicles suspended in K+ buffer. The effect of FCCP + valinomycin is not sensitive to external pH values between 6.3 and 9.1 and is diminished at lower external pH. The possible role of intravesicular pH and of the proton electrochemical gradient in the storage of acetylcholine within cholinergic vesicles is discussed.  相似文献   

17.
Micropuncture and microanalytical methods were employed to investigate the rôle of the spermathecal epithelium of the honey queen-bee in providing the appropriate conditions for the prolonged storage of spermatozoa. It was found that the epithelium maintains large concentration gradients of inorganic ions, generates an electrical potential difference of 21 mV, lumen positive, and produces a pH difference of up to 2.4 pH units between spermathecal fluid (SF) and haemolymph (H). The SFH concentration ratios for K+, Na+, Ca++, Cl?1, HPO4??, H2PO4? and amino acids were 7.7; 0.5; 0.8; 0.4; 1.03; 0.004; 0.3, respectively. While the pH value of haemolymph was constant at 6.17, the pH of SF increased with age from 7.3 to 8.6 over the first 3 days. The calculated electrochemical potential differences suggest that the epithelium of the spermathecal wall secretes K+ (and possibly HCO3? or OH?) actively into the lumen, but handles Na+ passively. This pattern conforms with the organization of ion transport in other insects.  相似文献   

18.
(1) Putative relationships between the rate of photophosphorylation, the proton-motive force and the concentration of an uncoupling molecule are considered within the framework of the delocalised chemiosmotic coupling hypothesis. The addition of a partially inhibitory titre of a specific, tight-binding H+-ATP synthase inhibitor is not expected, within the framework of a delocalised coupling model, to alter the form of this relationship. (2) Photophosphorylation in chromatophores from Rhodopseudomonas capsulata is potently uncoupled by the protonophore SF6847. Yet the uncoupling potency of this compound is actually further increased when the rate of phosphorylation in the absence of protonophore is decreased by the addition of the energy-transfer inhibitor venturicidin, in contrast to the expectations of a delocalised energy-coupling model. (3) Similarly, valinomycin (in the presence of nigericin) uncouples more potently when the number of active H+-ATP synthases is decreased by the addition of the energy -transfer inhibitors N,N′-dicyclohexyl-carbodiimide or venturicidin. (4) The pore-forming ionophore gramicidin D also uncouples photophosphorylation more potently when the number of active H+-ATP synthases is reduced. (5) These results are discussed in relation to the idea that the functional unit of electrical events and photophosphorylation either is, or is not, the intact membrane vesicle. (6) It is concluded that the unit of energy coupling in bacterial chromatophores is much smaller than the entire coupling membrane vesicle, and that previous analyses of this point, based on titrations with ionophores alone, may need to be re-examined.  相似文献   

19.
1. Both valinomycin and p-trifluoromethoxy carbonyl cyanide phenylhydrazone (FCCP) are required for full release of respiration by cytochrome c oxidase-containing proteoliposomes (prepared by sonicating beef heart cytochrome aa3 in salt solution with 4 parts phosphatidylcholine, 4 parts phosphatidylethanolamine and 2 parts cardiolipin) in the presence of external ascorbate and cytochrome c. In the absence of valinomycin the response to FCCP is rather sluggish, as reported by Wrigglesworth et al. (1976) (Abstracts, 10th Int. Congr. Biochem., No. 06-6-230).2. The Km for cytochrome c in 67 mM, pH 7.4, phosphate buffer with ascorbate as substrate, was 9 μM in both absence and presence of valinomycin and FCCP. Energization thus acts non-competitively towards cytochrome c oxidation.3. The apparent Km for oxygen is greater in the energized than in the deenergized state; double reciprocal plots of respiration rate versus oxygen concentration are concave downward in the absence of uncouplers, as found with intact mitochondria. Energization thus acts “competitively” towards oxygen.4. Despite the lack of a functional ATPase system, all the kinetic features of energization found in intact mitochondria can be mimicked in the reconstituted liposomes. This supports the chemiosmotic idea that electrical and perhaps H+ gradients modify the oxidase activity in reconstituted vesicles.  相似文献   

20.
We have previously shown that anacardic acid has an uncoupling effect on oxidative phosphorylation in rat liver mitochondria using succinate as a substrate (Life Sci. 66 (2000) 229-234). In the present study, for clarification of the physicochemical characteristics of anacardic acid, we used a cyanine dye (DiS-C3(5)) and 9-aminoacridine (9-AA) to determine changes of membrane potential (ΔΨ) and pH difference (ΔpH), respectively, in a liposome suspension in response to the addition of anacardic acid to the suspension. The anacardic acid quenched DiS-C3(5) fluorescence at concentrations higher than 300 nM, with the degree of quenching being dependent on the log concentration of the acid. Furthermore, the K+ diffusion potential generated by the addition of valinomycin to the suspension decreased for each increase in anacardic acid concentration used over 300 nM, but the sum of the anacardic acid- and valinomycin-mediated quenching was additively increasing. This indicates that the anacardic acid-mediated quenching was not due simply to increments in the K+ permeability of the membrane. Addition of anacardic acid in the micromolar range to the liposomes with ΔΨ formed by valinomycin-K+ did not significantly alter 9-AA fluorescence, but unexpectedly dissipated ΔΨ. The ΔΨ preformed by valinomycin-K+ decreased gradually following the addition of increasing concentrations of anacardic acid. The ΔΨ dissipation rate was dependent on the pre-existing magnitude of ΔΨ, and was correlated with the logarithmic concentration of anacardic acid. Furthermore, the initial rate of ΔpH dissipation increased with logarithmic increases in anacardic acid concentration. These results provide the evidence for a unique function of anacardic acid, dissimilar to carbonylcyanide p-trifluoromethoxyphenylhydrazone or valinomycin, in that anacardic acid behaves as both an electrogenic (negative) charge carrier driven by ΔΨ, and a ‘proton carrier’ that dissipates the transmembrane proton gradient formed.  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