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.
The dependence of the surface potential difference (ΔU), transversal elasticity module (E1) and membrane conductivity (G0) on the concentrations of the antiviral drugs, rimantadine and amantadine was studied in the planar bilayer lipid membrane system. The method used was based on independent measurements of the second and third harmonics of the membrane capacitance current. The binding constants of bilayer lipid membranes obtained from the drug adsorption isotherms were 2.1 · 105 M?1 and 1.3 · 104 M?1 for rimantadine and amantadine, respectively. The changes in G0 took place only after drug adsorption saturation had been achieved. The influence of rimantadine and amantadine on the interaction of bilayer lipid membranes with matrix protein from influenza virus was also investigated. The presence of 70 μg/ml rimantadine in the bathing solution resulted in an increase in the concentration of M-protein at which the adsorption and conductance changes were observed. The effects of amantadine were similar to those of rimantadine but required a higher critical concentration of amantadine. The results obtained suggest that the antiviral properties of rimantadine and amantadine may be related to the interaction of these drugs with the cell membrane, which can affect virus penetration into the cell as well as maturation of the viral particle at the cell membrane.  相似文献   

2.
Uptake of 14C-labelled sucrose and glucose by isolated seed coat halves of pea (Pisum sativum L. cv. Marzia) seeds was measured in the concentration range <0.1 μM to 100 mM. The initial influx of sucrose was strictly proportional to the external concentration, with a coefficient of proportionality (k) of 6.2 μmol·(g FW)?1·min?1·M?1. Sucrose influx was not affected by 10 μM carbonylcyanide m-chlorophenylhydrazone (CCCP), but it was inhibited by 40% in the presence of 2.5 mM p-chloromercuribenzenesulfonic acid (PCMBS). Influx with diffusional kinetics was also observed for glucose (k = 4.8 μmol·(g FW)?1·min ?1·M ?1) and mannitol (k = 5.1 μmol·(g FW)?1·min?1·M?1). For glucose an additional saturable system was found (Km = 0.26 mM, V max = 4.2 nmol·(g FW)?1·min?1), which appeared to be completely inhibited by CCCP and partly by PCMBS. In contrast to the diffusional pathway, uptake by this saturable system was slightly pH-dependent, with an optimum at pH 5.5. The influx of sucrose appears to be by the same pathway as the efflux of endogenous sucrose, which was inhibited by 36% in the presence of 2.5 mM PCMBS (De Jong A, Wolswinkel P, 1995, Physiol Plant 94: 78–86). It is argued that passive transport may be the only mechanism for sucrose transport through the plasma membrane of seed coat parenchyma cells. The estimated permeability coefficient of the plasma membrane for sucrose (P = 3.5·10?7 cm·s?1) is more than 1 × 106-fold higher than that reported for artificial lipid membranes. This relatively high permeability is hypothesized to result from pore-forming proteins that allow the diffusion of sucrose. Furthermore, it is shown that a sucrose gradient across the plasma membrane of the seed coat parenchyma of only 22 mM will suffice to result in the net efflux of sucrose which is required to feed the embryo.  相似文献   

3.
4.
Single spherical bilayer membranes of the Pagano-Thompson type (Pagano, R. and Thompson, T.E. (1967) Biochim. Biophys. Acta 144, 666–669), formed from monooleyl phosphate and cholesterol dissolved in CHCl3/CH3OH/n-decane, were subjected to a fast impedance analysis of high precision. Dielectric behavior of the whole system, as monitored from outside the spherical membrane, was sensitive to changes in the membrane state from the thick colored to the thin black state. With a spherical membrane 2–3 mm in diameter formed in the sample cavity containing 0.12 ml 10 mM NaCl, the former state was characterized by a dielectric dispersion having dielectric increment (Δ?) of some 102 and characteristic frequency (?c) around 106 Hz, while the latter had Δ? ? 105and ?c ? 103Hz. Complex plane plots for both dispersions traced semicircles, indicating that the present system may be unequivocally analyzed to yield spherical radius and membrane capacity (Cm) on the basis of a well-established dielectric theory. Cm for the thin membranes has thus been determined to be 0.54 μF · cm?2, in excellent agreement with a separate determination on planar membranes. The applicability of the present type of spherical membranes under dielectric monitoring to the study of membrane fusion or of exocytosis is suggested.  相似文献   

5.
Diffusion of thiocyanate (SCN?) and thiocyanic acid (HSCN) (pK=?1.8) through lipid bilayer membranes was studied as a function of pH. Membranes were made of egg phosphatidylcholine or phosphatidylcholine plus cholesterol (1:1 mol ratio) dissolved in decane or tetradecane. Tracer fluxes and electrical conductances were used to estimate the permeabilities to HSCN and SCN?. Over the pH range 1.0 to 3.3 only HSCN crosses the membrane at a significant rate. The relation between the total SCN flux (JA), concentrations and permeabilities is: 1/JA=1/Pul([A?]+[HA])+1/PHAm[HA], where [A?] and [HA] are the concentrations of SCN? and HSCN, Pul is permeability coefficient of the unstirred layer, and PHAm is the membrane permeability to HSCN. By fitting this equation to the data we find that PHAm = 2.6 cm · s?1 and Pul = 9.0 · 10?4 cm · s?1. Conductance measurements indicate that PA?m is 5 · 10?9 cm · s?1. Addition of cholesterol to phosphatidylcholine (1:1 mol ratio) reduces PHAm by a factor of 0.4 but has no effect on PA?m. SCN? is potent inhibitor of acid secretion in gastric mucosa, but the mechanism of SCN? action is unknown. Our results suggest that SCN? acts by combining with H+ in the mucosal unstirred layer (secretory pits) and diffusing back into the cells as HSCN, thus dissipating the proton gradient across the secretory membrane. A similar mechanism of action is proposed for some other inhibitors of gastric acid secretion, e.g. nitrite (NO2?), cyanate (CNO?) and NH4+.  相似文献   

6.
Abstract. Net efflux of Cl? was measured potentiometrically (Ag/AgCl electrode) during turgor regulation which was induced by hypotonic treatment (hypotonic turgor regulation) in the alga Lamprothamnium succinctum. The efflux of Cl? reached the peak value (11 μmol m ?2s?1) several minutes after the hypotonic treatment was started and then declined. The efflux of Cl? and inhibition of the cytoplasmic streaming [reflection of an increase in cytoplasmic concentration of free Ca2+([Ca2+]c)] were blocked under a low external concentration of Ca2+ ([Ca2+]e) (0·01 mol m?3) and resumed after raising [Ca2+]e to the normal value (3·9 mol m?3). The decrease in cell-osmotic pressure upon hypotonic treatment was inhibited by lowering either turgor pressure or [Ca2h]e. The inhibition was reflected in decreases of both the efflux of Cl? and the membrane conductance. Recovery of the cytoplasmic streaming from the inhibition was also accelerated by the same treatments. It is concluded that an increase in turgor pressure is continuously sensed by the cells and that continuous presence of external Ca2+ is necessary for the hypotonic turgor regulation.  相似文献   

7.
One of the major proteins of the outer membrane of Escherichia coli, the matrix protein (porin), has been isolated by detergent solubilisation. When the protein is added in concentrations of the order of 10 ng/cm3 to the outer phases of a planar lipid bilayer membrane, the membrane conductance increases by many orders of magnitude. At lower protein concentrations the conductance increases in a stepwise fashion, the single conductance increment being about 2 nS (1 nS = 10?9siemens = 10?9 Ω?1) in 1 M KCl. The conductance pathway has an ohmic current vs. voltage character and a poor selectivity for chloride and the alkali ions. These findings are consistent with the assumption that the protein forms large aqueous channels in the membrane. From the average value of the single-channel conductance a channel diameter of about 0.9 nm is estimated. This channel size is consistent with the sugar permeability which has been reported for lipid vesicles reconstituted in the presence of the protein.  相似文献   

8.
The combined action of electric field (105–107 V · m?1) and light (380–580 nm, 80 W · m?2) activating the photoenergetic reaction of bacteriorhodopsin (BR) in dry films of purple membranes from Halobacterium halobium was studied. A new stimulating effect of the field on the BR412 intermediate accumulation in the normal photochromic cycle of BR570 has been observed. The formation of the product BR412 is supposed to be accompanied by specific rearrangements of certain charged, polar and polarizable groups in the BR pigment-protein matrix. Such an intrinsic polarization could be promoted by an external electric field, the displacement vector of those groups being oriented in the direction of the field. The dielectric polarization properties of the purple membranes have been demonstrated by electret-thermal analysis.  相似文献   

9.
Abstract When internodal cells of Lamprothamnium succinetum were exposed to a hypotonic medium containing more than 1 mol m?3 Ca2+, the elevated turgor pressure decreased and reached a steady state within 30–60 min. The hypotonic treatment caused the membrane potential to depolarize, with a time lag of ca. 1 min. The membrane conductance increased transiently with the same time lag and reached a peak value within 2–3 min. When the external Ca2+ concentration was lowered to 0.01 mol m?3, both turgor regulation and change in the membrane conductance were strongly inhibited, although the membrane depolarization was not affected. When the Ca2+ level was returned to the normal level, the cells recovered their ability for turgor regulation and the membrane conductance attained a peak value after ca. 15–30 s. This response time is definitely shorter than that needed for the conductance change in cells exposed to a hypotonic medium having a normal level of Ca2+ from the beginning. We thus conclude that at least two sequential processes are involved in turgor regulation: a Ca2+ -independent process, followed by a Ca2+-dependent process.  相似文献   

10.
All-trans retinoyl fluoride was prepared by treating all-trans retinoic acid with diethylaminosulfurtrifluoride. The crystalline product, which was characterized by melting point, infrared, 1H-NMR, 19F-NMR and elementary analysis, showed λmax at 382 nm in hexane (ε = 4.98·104 M?1·cm?1) and at 392 nm in methanol (ε = 4.60·104 M?1·cm?1). Its biological activity in the rat growth assay, relative to all-trans retinyl acetate, was 22% ± 10%. Upon oral administration for 5 days to vitamin A-depleted rats, retinoyl fluoride (1020 μg) was rapidly metabolized to a polar metabolite fraction and, in the intestine, to an unstable retinol-like metabolite, purpotedly 15-fluororetinol. Upon administering intraperitoneally smaller doses (47–94 μg) of [11-3H]retinoyl fluoride, which was synthesized from [11-3H] retinoic acid, radioactive retinoic acid was noted in the liver and plasma but not in the intestine. As expected, a radioactive polar fraction appeared in the intestine and liver, but radioactive retinol, retinyl ester and some common oxidation products were not detected. Of the administered radioactivity, 72% was excreted in the urine, and only 4% was found in the feces over a 7-day period. Hydrolysis of the urine gave a radioactive fraction with a polarity similar to that of retinoic acid. Retinoyl fluoride also reacts readily with glycine to yield N-retinoyl glycine. Thus, the biological activity of retinoyl fluoride can be attributed to the formation of retinoic acid, probably by way of N-retinoyl derivatives. A possible pathway for its metabolism is presented.  相似文献   

11.
The small intestine is known to possess mechanisms for intact transport and membrane hydrolysis of oligopeptides. To determine the relative role of these processes in peptide assimilation the fate of two model peptides known to be high-affinity substrates for the brush border aminooligopeptidase were studied in rat small intestine in vivo. Both 20 mM Gly-L-Pro, a potent inhibitor of peptide transport, and specific inhibitors of the aminopeptidase, 10 mM L-Ala-β-naphthylamide or the phthalimido derivative of 0.1 mM L-leucine bromomethyl ketone, reduced assimilation of L-Leu-Gly-Gly and L-Leu-L-Leu. Further inhibition was found when both transport and peptidase inhibitors were included in the intestinal perfusate suggesting that the model di- and tripeptides utilize both intact transport and surface hydrolysis for their assimilation. Although comparative kinetic parameters of intact transport (Km = 22 mM; V = 1.9 · 10?3μmol · s?1 · cm?2) and surface hydrolysis (Km = 8.7; V = 1.1 · 10?3) for l-Leu-l-Leu differed markedly, the relationship of peptide concentration to assimilation rate was nearly identical for intact transport and surface hydrolysis in the physiological range of 1–10 mM substrate. Both intact peptide transport and surface hydrolysis appear to be efficient and complementary processes that promote efficient assimilation of dipeptides and tripeptides. The relative importance of each assimilation process appears to depend upon the amino acid composition of the peptide nutrient.  相似文献   

12.
The resealing process of lipid bilayer membranes after reversible electrical breakdown was investigated using two voltage pulses switched on together. Electrical breakdown of the membranes was induced with a voltage pulse of high intensity and short duration. The time course of the change in membrane conductance after the application of the high (short) voltage pulse was measured with a longer voltage pulse of low amplitude. The decrease in membrane conductance during the resealing process could be fitted to a single exponential curve with a time constant of 10-2 μs in the temperature range between 2 and 20°C. The activation energy for this exponential decay process was found to be about 50 kJ/mol, which might indicate a diffusion process. Above 25°C the resealing process is controlled by two exponential processes.The data obtained for the time course of the resealing process can be explained in terms of pore formation in the membranes in response to the high electrical field strength. A radius of about 4 nm is calculated for the initial pore size. From the assumed exponential change of the pore area with progressive resealing time a diffusion constant of 10?8 cm2/s for lateral lipid diffusion can be estimated.  相似文献   

13.
Kinetic study of the reaction between tocopheroxyl (vitamin E radical) and reduced ubiquinone, n = 10) has been performed. The rates of reaction of ubiquinol with α-tocopheroxyl 1 and seven kinds of alkyl substituted tocopheroxyl radicals 2–8 in solution have been determined spectrophotometrically, using a stopped-flow technique. The result shows that the rate constants decrease as the total electron-donating capacity of the alkyl substituents on the aromatic ring of tocopheroxyls increases. For the tocopheroxyls with two alkyl substituents at ortho positions (C-5 and C-7), the second-order rate constants, k1, obtained vary i n the order of 102, and decrease predominantly, as the size of two ortho-alkyl groups (methyl, ethyl, isopropyl and tert-buty) in tocopheroxyl increases. On the other hand, the reaction between tocopheroxyl and ubiquinone-10 (oxidized ubiquinone) has not been observed. The result indicates that ubiquinol-10 regenerates tocopherol by donating a hydrogen atom of the 1-OH and/or 4-OH group to the tocopheroxyl radical. For instance, the k1 values obtained for α-tocopheroxyl are 3.74 · 105 M?1 · s?1 and 2.15 · 5 M?1 · s?1 in benzene and ethanol solution at 25°C, respectively. The above reaction rates, k1, obtained were compared with those of vitamin C with α-tocopheroxyl reported by Packer et al. (k2 = 1.55 · 106 M?1 · s?1) and Scarpa et al. (K2 = 2 · 105 105 M?1 · s?1), which is well known as a usual regeneration reaction of tocopheroxyl in biomembrane systems. The result suggests that ubiquinol-10 also regenerates the tocopheroxyl to tocopherol and prevents lipid peroxidation in various tissues and mitochondria.  相似文献   

14.
The uptake and efflux of 22Na and 42K were studied in denuded Rana pipiens oocytes following progesterone induction of the resumption of meiotic maturation. Coincident with the breakdown of the large nucleus, or germinal vesicle, there is a virtual disappearance of K+ permeability of the oocyte plasma membrane. Only about 1–2% of the total [K+]i is exchanged by completion of nuclear breakdown (8–10 hr) and accounts for the finding that there is no detectable change in total [K+]i during the first meiotic division (20–24 hr). In the case of Na+, influx, exchange, and efflux kinetics were unchanged during the first meiotic division, with 20 and 35% of the total oocyte Na+ exchanging by the completion of nuclear breakdown and first meiotic division, respectively. Removal of Na+ from the incubation medium produced and earlier nuclear breakdown, whereas a K-free medium delayed breakdown. There was no effect of 10 μm/ml tetrodotoxin or 10?5M strophanthidin on the time course of nuclear breakdown. Thus one action of progesterone appears to be a selective turning off of “K channels” in the oocyte plasma membrane. The disappearance of K selectivity of the oocyte plasma membrane coincides with plasma membrane depolarization, as well as nuclear swelling and breakdown.  相似文献   

15.
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   

16.
Xylem probe measurements in the roots of intact plants of wheat and barley revealed that the xylem pressure decreased rapidly when the roots were subjected to osmotic stress (NaCl or sucrose). The magnitude of the xylem pressure response and, in turn, that of the radial reflection coefficients (σr) depended on the transpiration rate. Under very low transpiration conditions (darkness and high relative humidity), σr assumed values of the order of about 0·2–0·4. The σr values of excised roots were also found to be rather low, in agreement with data obtained using the root pressure probe of Steudle. For transpiring plants (light intensities at least 10 μmol m?2 s?1; relative humidity 20–40%) the response was nearly 1:1, corresponding to radial reflection coefficients of σr= 1. Further increase of the light intensity to about 400 μmol m?2 s?1 resulted in a slight but significant decrease of the σr values to about 0·8. Similar measurements on maize roots confirmed our previous results (Zhu et al. 1995, Plant, Cell and Environment 18, 906–912) that, in intact transpiring plants at low light intensities of about 10 μmol m?2 s?1 and at relative humidities of 20–40% as well as in excised roots, the xylem pressure response was much less than expected from the external osmotic pressure (σr values 0·3–0·5). In contrast to wheat and barley, very high light intensities (about 700 μmol m?2 s?1) were needed to shift the radial reflection coefficients of maize roots to values of about 0·9. Osmotically induced xylem pressure changes were apparently linked to changes in turgor pressure in the root cortical parenchyma cells, as shown by simultaneous measurements of xylem and cell turgor pressure. In analogy to the σr values of the respective glycophytes, the σc values of the root cortical cells of wheat and barley were close to unity, whereas σc for maize was significantly smaller (about 0·7) under laboratory conditions. When the light intensity was increased up to about 700 μmol m?2 s?1 the cellular reflection coefficient of maize roots increased to about 0·95. In contrast to the σr values, the σc values of the three species investigated remained almost unchanged when the leaves were exposed to darkness and humidified air or when the roots were cut. The transpiration-dependent (species-specific) pattern of the cellular and radial reflection coefficients of the root compartment of the three glycophytes apparently resulted from (flow-dependent) concentration-polarization and sweep-away effects in the roots of intact plants. The data could be explained straightforwardly terms of theoretical considerations outlined previously by Dainty (1985, Acta Horticulturae 171, 21–31). The far-reaching consequences of this finding for root pressure probe measurements on excised roots, for the occurrence of pressure gradients under transpiring conditions, and for the non-linear flow-force relationships in roots found by other investigators are discussed.  相似文献   

17.
E. Steudle  W. D. Jeschke 《Planta》1983,158(3):237-248
Radial transport of water in excised barley (Hordeum distichon, cv. Villa) roots was measured using a new method based on the pressure-probe technique. After attaching excised roots to the probe, root pressures of 0.9 to 2.9 bar were developed. They could be altered either by changing the root pressure artificially (with the aid of the probe) or by changing the osmotic pressure of the medium in order to induce water flows across the root. The hydraulic conductivity of the barley roots (per cm2 of outer root surface) was obtained in different types of experiments (initial water flow, pressure relaxations, constant water flow) and was (0.3–4.3)·10-7 cm s-1 bar-1. The hydraulic conductivity of the root was by an order of magnitude smaller than the hydraulic conductivity of the cell membranes of cortical and epidermal cells (0.8–2.2)·10-6 cm s-1 bar-1. The half-times of water exchange of these cells was 1–21 s and two orders of magnitude smaller than that of entire excised roots (100–770 s). Their volumetric elastic modulus was 15–305 bar and increased with increasing turgor. Within the root cortex, turgor was independent of the position of the cell within a certain layer and turgor ranged between 3 and 5 bar. The large difference between the hydraulic conductivity of the root and that of the cell membranes indicates that there is substantial cell-to-cell (transcellular plus symplasmic) transport of water in the root. When it is assumed that 10–12 membrane layers (plasmalemma plus tonoplast) in the epidermis, cortex and endodermis form the hydraulic resistance to water flow, a value for the hydraulic conductivity of the root can be calculated which is similar to the measured value. This picture for water transport in the root contradicts current models which favour apoplasmic water transport in the cortex.  相似文献   

18.
The effect of synthetic anti-oxidant potassium phenosan (PP, potassium salt of β-(4-hydroxy-3,5-ditretbutil-phenyl)-propionic acid) on the structural state of the surface (8 Å) and deep (20–22 Å) lipid regions of plasma membranes of mice liver cells was studied by spin probes method in vitro in a wide range of concentrations (10?5–10?21 M). Two stable free radicals, 5- and 16-doxyl-stearic acids (C5 and C16), were used as spin probes. The nonlinear polymodal dose-effect dependences were obtained for parameters that characterize the microviscosity of the lipid bilayer (τc) in the site of localization of the probe C16, and the order parameter (S), which characterizes the stiffness of the surface layers of lipids in the site of localization of the probe C5. Statistically a reliable increase was observed for parameter τc after addition of PP at concentrations 10?5–10?7 M and 10?18–10?19 M, and for parameter S after addition of PP at concentrations 10?6–10?7 M and 10?13–10?15 M. Peaks on both dose-effect curves were separated by the intervals of concentrations where PP had no effect on the studied physico-chemical characteristics of biomembranes. For PP concentrations which caused maximal changes in τc and S, we investigated thermal dependence of these parameters and determined the thermally induced structural transitions. Comparing with control, ultra-low doses of PP (10?13–10?15 M) and (10?18–10?19 M) caused an appearance of additional thermally induced structural transition in the surface and deep regions of plasma membrane lipids. The possible role of the interaction of PP molecules with specific binding sites on plasma membranes and formation of nanoparticles of PP in very dilute aqueous solutions are discussed.  相似文献   

19.
Small amphiphilic compounds (Mr<200 Da) such as anaesthetics and hexane derivatives with different polar groups produced a concentration-dependent acceleration of the slow passive transbilayer movement of NBD-labelled phosphatidylcholine in the human erythrocyte membrane. Above a threshold concentration characteristic for each compound, the flip rate gradually increased at increasing concentrations in the medium. For compound concentrations required to produce a defined flip acceleration, corresponding membrane concentrations were estimated using reported octanol/water partition coefficients. The effective threshold membrane concentrations (50–150 mmol l?1) varied in the order: hexylamine>isoflurane=hexanoic acid>hexanol=chloroform>hexanethiol=1,1,2,2-tetrachloroethane>chlorohexane. Apolar hexane, which mainly distributes in the apolar membrane core, was much less effective and supersaturating concentrations were required to enhance flip. Localization of the drug at the lipid–water interface seems to be required for flip acceleration. Such a localization may increase the lateral pressure in this region and the bilayer curvature stress with concomitant decrease of order and rigidity at the interface. This unspecific bilayer perturbation is proposed to enhance the probability of formation of hydrophobic defects in the bilayer, facilitating penetration of the polar head group of the phospholipid into the apolar membrane core.  相似文献   

20.
Most antimicrobial peptides exert their rapid bactericidal activity through a unique mechanism of bacterial membrane disruption. However, the molecular events that underlie this mechanism remain partly unresolved. In this study, the frequency shift (ΔF) obtained through quartz-crystal microbalance with dissipation (QCM–D) indicated that the initial binding of Ib-AMP4 within the lipid membrane started at a critical Ib-AMP4 concentration that exceeded 100 μg/ml. Circular dichroism measurements provided evidence that Ib-AMP4 occurs in a β-sheet configuration which is adapted for insertion into the lipid membrane. Monolayer experiments and the value of dissipation alteration (ΔD) obtained through QCM–D showed that the pressure increased within the phospholipid bilayer upon peptide insertion, and the increase in pressure subsequently forced the bilayer to wrinkle and form pores. However, D continued to increase, indicating that the membrane surface underwent a dramatic morphological transition: the membrane surface likely became porous and uneven as Ib-AMP4 projected from the external surface of the lipid bilayer. Intensive peptide insertion, however, soon plateaued 1 min after the addition of Ib-AMP4. This behaviour corresponded with the results of bactericidal kinetics and liposome leakage assays. A sudden decrease in D accompanied by a negligible decrease in F occurred after replacing the Ib-AMP4 solution with HEPES buffer. This result implied that the bilayer surface rearranged and that poration and wrinkling decreased without further peptide insertion. Transmission electron microscopy results indicated that pore formation occurred during Ib-AMP4 insertion but eventually subsided. Therefore, the mode of action of AMP in bacterial membranes could be elucidated through QCM–D.  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