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 interpretation of electrical breakdown in terms of electro-mechanical instabilities, predicts that the breakdown potential should decrease with increasing cell turgor pressure.Experiments were conducted to test this hypothesis on cells of Valonia utricularis over a turgor pressure range of 0.5 · 105–5.0 · 105 N/m2. Electrical breakdown was measured using intracellular electrodes and 500 μs current pulses. The pressure was monitored by an intracellular micropipette pressure transducer. The results obtained show a linear decrease in the critical breakdown potential with pressure. The effective compressive modulus of the cell membrane, γ, is calculated from the slope of this line to 69 ± 10 · 105N/m2 (average value of seven measurements). This is consistent with the theoretical prediction of the electromechanical model using our previously determined values of the elastic modulus of the membrane.A theoretical analysis is given of the effects of pressure on the breakdown. This includes also considerations of the indirect effect of pressure on the membrane via stretching of the cell wall with a possible coupling of such strains to the cell membrane. The results and analysis presented allow us to conclude on the basis of the experimentally determined breakdown P.D. of 959 mV that the region of membrane where electrical breakdown occurs is a dielectric with one of the following combinations of parameters: (A) a thickness δ = 7–9 nm with a dielectric constant ? = >10, e.g. a hydrated protein spanning the whole membrane. (B) δ = 4–5 nm with ? = 3–8, e.g. a lipoprotein of lipid bilayer dimensions. (C) δ ≈ 2 nm with ? = 2–3, e.g. a half lipid bilayer.If we assume that the breakdown P.D. of the tonoplast and plasmalemma are identical, that is 480 mV, then there is only one reasonable choice for the membrane thickness and the dielectric constant: δ = 2nm, ? = 3–8, e.g. a (lipo-)proteinaceous module facing a half lipid bilayer.  相似文献   

2.
3.
Membrane surface potential on the periplasmic side of the photosynthetic membrane was estimated in cells, spheroplasts and chromatophores of Rhodopseudomonas sphaeroides. When the membrane potential (potential difference between bulk aqueous phases) was kept constant in the presence of carbonylcyanide m-chlorophenylhydrazone, addition of salt to a suspension of cells or spheroplasts induced a red shift in the carotenoid absorption spectrum which indicated a change in the intramembrane electrical field. The spectral shift is explained by a rise in electrical potential at the outside surface of the photosynthetic membrane due to a decrease in extent of the negative surface potential.The spectral shift occurred in the direction opposite to that in chromatophores, indicating that the sidedness of the membrane of cells or spheroplasts is opposite to that of chromatophores. The dependences of the extent of the potential change on concentration and valence of cations of salts agreed with the Gouy-Chapman relationship on the electrical diffuse double layer. The charge density on the periplasmic surface of the photosynthetic membrane was estimated to be ?2.9 · 10?3 elementary charge per Å2, while that on the cytoplasmic side surface was calculated as ?1.9 · 10?3 elementary charge per Å2 (Matsuura, K., Masamoto, K., Itoh, S. and Nishimura, M. (1979) Biochim. Biophys. Acta 547, 91–102). Surface potential on the periplasmic side of the photosynthetic membrane was estimated to be about ?50 mV at pH 7.8 in the presence of 0.1 M monovalent salt.  相似文献   

4.
Electron transfer between horse heart and Candida krusei cytochromes c in the free and phosvitin-bound states was examined by difference spectrum and stopped-flow methods. The difference spectra in the wavelength range of 540–560 nm demonstrated that electrons are exchangeable between the cytochromes c of the two species. The equilibrium constants of the electron transfer reaction for the free and phosvitin-bound forms, estimated from these difference spectra, were close to unity at 20°C in 20 mM Tris-HCl buffer (pH 7.4). The electron transfer rate for free cytochrome c was (2–3) · 104 M?1 · s?1 under the same conditions. The transfer rate for the bound form increased with increase in the binding ratio at ratios below half the maximum, and was almost constant at higher ratios up to the maximum. The maximum electron exchange rate was about 2 · 106 M?1 · s?1, which is 60–70 times that for the free form at a given concentration of cytochrome c. The activation energy of the reaction for the bound cytochrome c was equal to that for the free form, being about 10 kcal/mol. The dependence of the exchange rate on temperature, cytochrome c concentration and solvent viscosity suggests that enhancement of the electron transfer rate between cytochromes c on binding to phosvitin is due to increase in the collision frequency between cytochromes c concentrated on the phosvitin molecule.  相似文献   

5.
Adenosine 3′,5′-monophosphate (cyclic AMP) (10−5 g/ml) often caused a gradual increase in spotaneous contractile activity of the hindgut of the cockroach, Leucophaea maderae, and on rare occasions it would evoke a hormone-like response. However, aminophylline (2·5 × 10−4 g/ml) was capable of mimicking the neurohormone, and a concentration of 2·5 × 10−5 g/ml potentiated the contractile response evoked by the neurohormone: these responses were blocked by either the presence of 1 mM manganous ion or in a high potassium solution (162 mM). Propranolol (10−6 g/ml) and dopamine (10−4 g/ml) suppressed both spontaneous contractile events and neurohormone action. Dopamine (5 × 10−6 g/ml) also blocked action potential generation as did propranolol at 10−4 g/ml.These results lead us to suppose that cyclic AMP might serve as a mediator of neurohormone action by increasing calcium transport across the surface membrane of muscle fibres. Caffeine (2·5 × 10−4 g/ml), like aminophylline, caused a hormone-like response in normal hindguts. Even when the visceral muscles of the hindgut were depolarized in 162 mM potassium solution (without calcium), caffeine was still capable of inducing a phasic response. However, the addition of 2 mM calcium to such potassium-depolarized preparations caused a gradual increase in muscle tonus and substantially potentiated the response to caffeine.Such findings clearly implicate calcium as the mediator of excitation-contraction coupling in visceral muscle. While the interactions between the neurohormone, cyclic AMP, and calcium seem to be largely associated with the surface membrane and action potential generation.  相似文献   

6.
Reisin I.L., Rabito C.A. and Cantiello H.F. 1981. Water and electrolyte balance in protoscoleces of Echinococcus granulosus incubated in vitro: effect of metabolic inhibitors. International Journal for Parasitology 11: 405–410. The effects of metabolic inhibitors on the Na+, K+, Cl? and water balance of protoscoleces of Echinococcus granulosus (sheep strain) were studied in vitro. The protoscoleces were incubated at 37°C in Ringer Krebs solution for 3 h in the presence of iodoacetate, 3 mM (IA); potassium cyanide, 3 mm (KCN); 2?4 dinitrophenol, 0.2 mm (DNP); ouabain, 10?M or ethacrynic acid 0.5 mm. The effects of IA and/or KCN on the water and electrolyte balance were tested at high (0.95 × 105Pa) and low (0.05 × 105 Pa) oxygen tensions. Inhibitors produced a decrease in K+ as well as an increase in Na+ contents. At both high and low O2 tensions the Na+ balance was greatly altered by IA, the action of which could be already observed during the first hour of treatment. The cations did not reach a steady state balance during 3 h of incubation. At high oxygen tension Na+ and K+ balance was also altered by KCN or DNP though their actions were not as marked as that of IA. Ouabain affected the Na+ and K+ contents that reached new steady state distribution between 1.5 and 3 h of treatment while water and electrolyte contents were not modified by ethacrynic acid. In all the experiments no changes in Cl? and water contents were observed. It is concluded that the energy required to maintain the Na?K balance mechanisms within protoscoleces is largely provided by the anaerobic glycolytic pathway and that the aerobic oxidative pathway contribution to the energy balance is only accessory.  相似文献   

7.
8.
Synaptic vesicles from the Torpedo ray can be induced to aggregate in the presence of Ca2+ and K+ in the 4 mM and 50 mM range, respectively. The reactions are strikingly similar to those of chromaffin granule membranes reported previously (Morris, S.J., Chiu, V.C.K. and Haynes, D.H. (1979) Membrane Biochem. 2, 163–202). The Ca2+-induced reaction includes dimerization and higher order aggregation, and is shown to be due to electrostatic screening interactions and binding to negatively-charged groups on the membrane surface. The K+-induced reaction includes only dimerization and is shown to be due to screening interactions alone.The kinetics of the dimerization reactions were studied using the stopped-flow rapid mixing technique. The Ca2+-induced reaction has a ‘bimolecular’ rate constant of 4.77 · 108 M?1 · s?1 while the value for the K+-induced reaction is 7.05 · 109 M?1 · s?1. These values are close to the limit of diffusion control (8.03 · 109 M?1 · s?1), indicating that no large energy barriers or structural barriers to aggregation exist. Arrhenius plots for the Ca2+-induced aggregation showed a break at 5°C. Above this temperature, the activation energy is low (+0.65 kcal/mol), consistent with the above. Below this temperature, the activation energy is high, consistent with a membrane structure change increasing the energetic and structural barriers. This information, and the observation of a high stability constant of the complex, were taken as evidence for the involvement of ‘recognition sites’ on the membrane surface.  相似文献   

9.
J.A. Van Best  L.N.M. Duysens 《BBA》1975,408(2):154-163
The kinetics of the fluorescence yield Ф of chlorophyll a in Chlorella pyrenoidosa were studied under anaerobic conditions in the time range from 50 μs to several minutes after short (t12 = 30 ns or 5 μs) saturating flashes. The fluorescence yield “in the dark” increased from Ф = 1 at the beginning to Ф ≈ 5 in about 3 h when single flashes separated by dark intervals of about 3 min were given.After one saturating flash, Ф increased to a maximum value (4–5) at 50 μs, then Ф decreased to about 3 with a half time of about 10 ms and to the initial value with a half time of about 2 s. When two flashes separated by 0.2 s were given, the first phase of the decrease after the second flash occurred within 2 ms. After one flash given at high initial fluorescence yield, the 10-ms decay was followed by a 10 s increase to the initial value. After the two flashes 0.2 s apart, the rapid decay was not follewed by a slow increase.These and other experiments provided additional evidence for and extend an earlier hypothesis concerning the acceptor complex of Photosystem II (Bouges-Bocquet, B. (1973) Biochim. Biophys. Acta 314, 250–256; Velthuys, B. R. and Amesz, J. (1974) Biochim. Biophys. Acta 333, 85–94): reaction center 2 contains an acceptor complex QR consisting of an electron-transferring primary acceptor molecule Q, and a secondary electron acceptor R, which can accept two electrons in succession, but transfers two electrons simultaneously to a molecule of the tertiary acceptor pool, containing plastoquinone (A). Furthermore, the kinetics indicate that 2 reactions centers of System I, excited by a short flash, cooperate directly or indirectly in oxidizing a plastohydroquinone molecule (A2?). If initially all components between photoreaction 1 and 2 are in the reduced state the following sequence of reactions occurs after a flash has oxidised A2? via System I: Q?R2? + A → Q?R + A2? → QR? + A2?. During anaerobiosis two slow reactions manifest themselves: the reduction of R (and A) within 1 s, presumably by an endogenous electron donor D1, and the reduction of Q in about 10 s when R is in the state R? and A in the state A2?. An endogenous electron donor, D2, and Q? compete in reducing the photooxidized donor complex of System II in reactions with half times of the order of 1 s.  相似文献   

10.
Green thallus cells of the aquatic liverwort, Riccia fluitans, are rapidly depolarized in the presence of 1–20 μM NH4Cl and 5–100 μM CH3NH3Cl, respectively. Simultaneously, the membrane conductance is increased from 0.41 to 1.2 S · m?2. Uptake of [14C]methylamine is stimulated by increasing [K+]o and inhibited by increasing [Na+]o or [H+]o, is highly voltage sensitive, and saturates at low amine concentrations.Double-reciprocal plots of (a) maximal membrane depolarization and (b) methylamine uptake vs. external amine concentration give apparent Km values of 2 ± 1 μM ammonia and 25–50 μM methylamine; Km values for changes in conductance and membrane current are greater and voltage dependent. Whereas the amine transport into the cell is strongly inhibited by CN?, the amine efflux is stimulated.The current-voltage characteristics of the ammonia transport are represented by a sigmoid curve with an equilibrium potential of ?60 mV, and this is understood as a typical carrier curve with a saturation current of about 70 mA · m?2. It is further concluded that the evidently carrier-mediated transport is competitive for the two amines tested, and that ammonia and methylamine are transported in the protonated form as NH4+ and CH3NH3+ into the cytoplasm.  相似文献   

11.
Ambient sea-water nitrate and tissue nitrogen (ethanol soluble nitrate and amino acids, as well as total nitrogen) of Macrocystis integrifolia Bory were monitored over a 2-yr period in Bamfield, Vancouver Island, British Columbia. Sea-water nitrate varied from a high of 12 μmol · 1?1 (individual values as high as 23 μmol · 1?1 were recorded) in late winter to below detection limits for most of the summer. Tissue nitrate and total nitrogen paralleled the ambient nitrate levels and showed summer minima and winter maxima (from 0 to 70 μmol · g fresh wt?1 for nitrate and from 0.8 to 2.9% of dry wt for total N). The nitrate uptake capacity was inversely proportional to tissue nitrate concentration and, furthermore, was much higher for subapical surface blades (60–70 nmol · cm?2 · h?1) than for older, deeper blades (5–10 nmol · cm?2 · h?1). Nitrate uptake by subapical blade disks in summer is apparently higher in dark (1.0–1.7 μmol · g fresh wt?1 · h?1) than in light (0.6–1.3 μmol · g fresh wt?1 · h?1) and the data obtained in 36–108 h experiments indicate nitrate pool sizes of 30–90 μmol · g fresh wt?1. These pools are 23 to nearly full in winter. Ammonium does not inhibit nitrate uptake. It is taken up and apparently utilized much faster than nitrate and it may well be an important source of nitrogen for marine macrophytes.  相似文献   

12.
Twenty-one monthly collections of the Antarctic limpet Nacella (Patinigera) concinna (Strebel, 1908) were obtained by divers at Signy Island, South Orkney Islands. A mean monthly population density of 123.7 ± 21.2 · m?2, mean biomass of 13.7 ± 2.7 g dry tissue wt · m?2, and annual production of 2.9 g · m?2 were recorded in the depth range 2–12 m below mean low water.Shell growth was slow with a maximum growth rate, in the first 3–5 yr of life, of 3 4 mm per year. Maturity was attained at 7–8 yr (21 mm length), and maximum size (41 mm length) at about 21 yr. Unique spawning behaviour was observed in two Austral springs, and data relating spawning to the spring increase in sea temperature were obtained.  相似文献   

13.
An electrical potential having a value of ?35 mV (SD = 7.4) is recorded upon penetration of the ventral surface of adult male Schistosoma mansoni with a microelectrode. Histological studies using horseradish peroxidase as a marker injected iontophoretically through the recording electrode indicate that this potential originates in the tegumental epithelium. Recordings from animals with contractile activity showed spontaneous nonovershooting depolarizations with durations of 20 to 100 msec and amplitudes between 5 and 15 mV. Slow-wave depolarizations are also observed and are most frequent in animals showing only slight movement. The tegumental membrane potential appears to be dependent primarily on the K+ gradient across the surface since a 10-fold increase in external K+ causes a 30-mV change in the potential. Altering the external concentration of Cl? has little effect on the tegumental potential, but lowering the external Na+ concentration to 37 mM causes a 10-mV depolarization. Praziquantel (PZ) has little initial effect on the tegumental membrane potential. A slow depolarization does occur, however, which reaches a significant level about 10 min after PZ (1 μM). Carbachol and dopamine are without significant effects on the tegumental membrane potential.  相似文献   

14.
The action of chlordimeform on the nerve-muscle preparation of the larvae of the waxmoth Galleria mellonella has been studied by means of microelectrodes. Exitatory junction potential evoked by nerve stimulation is reversibly suppressed by 2 × 10?3 M chlordimeform, and spike-like component is abolished. The resting membrane potential of the muscle fibre and the action potential from the nerve terminal are not affected at 5 × 10?3 M chlordimeform. The depolarizing membrane response caused by outward current and the effective membrane resistance are not appreciably affected. It appears that chlordimeform exerts its blocking action on the neuromuscular junction rather than the conductance mechanism of muscle fibre membrane.  相似文献   

15.
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   

16.
The binding of cis(c)- and trans(t)-Pt(NH3)2Cl2 to DNA at platinum/DNA-nucleotide ratios (Ri) of 0.1 or less has been studied by means of radioactive 195mPt-labeled compounds. Kinetic data are consistent with the following scheme:
At 25°C and pH 5–6 in 5 mM NaClO4, the values for the rate constants in the above scheme for the c-isomer are k2 = 2.2 × 10?5 sec?1, k7 = 0.32 (sec M)?1, and k8 = 143 (sec M)?1; for the t-isomer the values are k2 < 0.5 × 10?5 sec?1 and k7 = 0.95 (sec M)?1. Platinum-DNA adducts do not undergo detectable exchange after 3 days at 37°C, indicating the absence of a dynamic equillibrium. For both isomers the rate of binding is the same for single- and double-stranded DNA. The conclusions derived from Ag+ and H+ titration studies are consistent with binding at guanine N(7) for Ri < 0.1. The reaction rate is competitively inhibited by various salts and buffers and is suppressed by raising the pH (50% inhibition of initial rates at pH 7.3). At 37°C and pH 7 in 0.15 M NaCl, 6–8% of both the c- and t-isomers bind to DNA in 24 h, suggesting that both compounds should bind to DNA under biological conditions.  相似文献   

17.
18.
Conventional hydrogen ion glas electrodes have been used for the preparation of enzyme-pH-electrodes by either entrapping the enzymes within polyacrylamide gels around the electrode or as liquid layer trapped within a cellophane membrane. The enzymes were glucose oxidase, urase and penicillinase.The pH response to glucose concentration was about linear within 10?1–10?3 M glucose and for urea linear within 5·10t—–5·10?5M. The pH response to penicillin was about linear in the range from 10?3–10?2 M resulting in a pH shift of 1.4 units; reproduceable pH response was obtained down to concentrations of 3·10?5 M.Studies as to the effect of buffer using an urease–pH-electrode showed a buffer concentration of 10?2 M a substantial shift of about one pH-unit in the range of 10?4 to 10?2 M urea. Both urease- and penicillinase–pH-electrodes were tested as to stability showing no decrease in pH response except at high substrate concentration (1·10?2 M) over a period of 2–3 weeks kept at room temperature.  相似文献   

19.
Salivary glands from third instar larvae of Drosophila melanogaster were incubated in vitro with various substances affecting oxidative phosphorylation. After an incubation time of 1–3 h changes in puff size and in cellular ATP level were registered. 10?6 M trinactin, 10?5 or 10?4 M oligomycin both induce puff 63BC together with some other puffs and reduce the cellular ATP level by about 80–90%. The trinactin-dependent puff induction can be inhibited, if the medium is supplemented with 10?3 M ATP or 10?3 M ITP or 10?6 M antimycin or 10?2 M KCN. The effect of exogenous ATP is prevented by adding 10?6 M oligomycin to the incubation mixture; 10?6 M oligomycin alone, however, has no inductive effect on 63BC. The presence of exogenous ITP, furthermore, prevents the ATP level from being reduced by trinactin. 10?4 M atractyloside lowers the ATP level by about 75 %, whereas a puff induction cannot be observed. The same is true for various concentrations of KCN. It is concluded that ATP itself is not involved directly in the regulation of puff activity but that it acts on a phosphorylating reaction that can be inhibited by oligomycin.  相似文献   

20.
During the latter half of July, 1980, a bloom of Gymnodinium flavum Kofoid & Swezy caused water discoloration in La Jolla Bay, California. This naked dinoflagellate dominated the phytoplankton, numeri- cally and by volume, and was found in concentrations as high as 6.15 × 103 cells · ml?1. It was most abundant near the surface above a shallow (5–10 m), sharp thermocline and a nitracline at 10 m. Near the end of July, the depth of maximum phytoplankton abundance descended and water discoloration was no longer visible at the surface even though areal concentrations of G. flavum did not decrease. Concurrent with changes in the vertical distribution of the phytoplankton, warm, nutrient-depleted water moved into the area and nitrate availability in the upper 20 m of the water column was drastically reduced. Measure- ments of the chemical composition of the phytoplankton do not, however, indicate progressive nutrient stress during the period of environmental change. We conclude that shoaling of the thermocline and nitracline associated with apparent upwelling were conducive to development of the bloom and that advection of warmer water from offshore led to disappearance of yellow surface water from the bay.  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