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 515 毫秒
1.
The proportionality constant between the equipotency concentrations of a series of solutes and the fraction of a solute in the membrane phase is directly related to the solute to lipid mol ratio. Experimental measurements of partition coefficient and of several alkanol-induced effects show that the solute/lipid mol ratlos for a series of alkanols are not constant at their equipotency concentrations. The deviations in the solute/lipid ratios are similar in the various systems, and these deviations seem to depend primarily upon the chain length and branching in alkanols. It is suggested that such intrinsic differences in the perturbing ability of alcohols arise from a specificity of interaction between alkanols and lipid bilayer. We have correlated partition coefficients (in n-octanol, in egg phosphatidylcholine liposomes, and in dipalmitoyl phosphatidylcholine liposomes) for thirteen alkanols to the equipotency concentrations for their ability to modify the order-disorder thermotropic transition in dipalmitoyl phosphatidylcholine, ability to stimulate the hydrolysis of phosphatidylcholine in a bilayer by bee venom phospholipase A2, and for the activation of the galactoside transport system in Escherichia coli. Significant correlation is found between equipotency concentrations for perturbing the order-disorder transition, the activation of phospholipase A2-catalyzed hydrolysis and the activation of galactoside transport system.  相似文献   

2.
Summary Ethanol, isopropanol, propanol and butanol exponentially inhibited the maximum velocity of the glucose transport system ofSaccharomyces cerevisiae, determined by use of the non-metabolizable analogued-xylose. While the exponential inhibition constants increased with the lipid solubility of the alkanols, they were independent of temperature in the range 21°–35°C: the Arrhenius plots (modified according to the theory of absolute reaction rates) of the initial maximum rates of xylose transport were linear and parallel in both the absence and presence of alkanols. Thus, the alkanols did not affect the enthalpy of activation of the glucose transport system (H ± was 12 190 cal mol-1) but decreased the entropy of activation. The following entropy coefficients (decrease in activation entropy per unit concentration of alkanol) were obtained: ethanol,-0.84; isopropanol,-1.21; propanol,-1.41 and butanol,-3.18 entropy units per mole per liter. The temperature relations of glucose fermentation with and without ethanol by resting cells over the temperature range studied (15°–35°C) were nearly identical with those of the glucose transport system, suggesting that the latter mediates the rate-limiting step of the former and that this relationship is maintained in the presence of ethanol.  相似文献   

3.
Summary The kinetic parameters for the steady-state rate of hydrolysis of egg phosphatidylcholine in multilamellar vesicles by bee venom phospholipase A2 are measured in the presence of 27 alkanols and several organic solvents. In general, small nonpolar solutes like enflurane, tetrahydrofuran, benzene, chloroform and diethylether do not promote the hydrolysis of multilamellar vesicles. The rate of hydrolysis shows a biphasic dependence upon the alkanol concentration for all higher (C5–C9) alcohols examined, i.e., an optimal rate of hydrolysis is observed at a characteristic concentration for each alcohol. The alkanol to lipid mole ratio (D/L ratio) in the bilayer at the peak activating concentration of an alkanol was computed from its bilayer/water partition coefficient. The branched chain alcohols induce peak activation of hydrolysis at lowerD/L ratios in the bilayer than the corresponding straight chain analogs. Similarly, the longer chainn-alkanols at peak activating concentration have a lowerD/L ratio than the corresponding lower alcohols. Both theK m andV m for phosphatidylcholine increase as a function of the chain length of the activating alcohol. These kinetic parameters also depend upon the position of the substituents on the activating alcohols. Both theD/L ratio andV m for an alcohol are found to change with the cross-sectional area of the activating alcohol across its long axis: alcohols with a more asymmetric cross-section exhibit higherV m and a lowerD/L ratio. Such correlations ofV m andD/L ratio with the molecular parameters of the alkanols are interpreted to suggest that the accessibility of the substrate molecule in the bilayer to the phospholipase is modulated by the free space introduced by the alkanols in the bilayer.Effect of tetradecane derivatives and A2C (a membrane fluidizing agent) on the phase transition characteristics of DPPC bilayers, and their susceptibility to phospholipase A2 from bee venom and pig pancreas is also reported. These solutes cause a broadening of the transition profile and reduce the size of the cooperative unit and the enthalpy of transition. These effects depend upon the mole fraction of a solute in the bilayer; however, equal concentrations of these solutes do not induce equal response. Susceptibility of the modified bilayers to phospholipase A2 depends not only upon the structure of the solute but also upon the source of the enzyme. The data show that the activity of the membrane-bound enzyme is modulated to different extents by different solutes, and the bilayer perturbing ability of these solutes may be related to the asymmetry of their cross-sectional area and to the free space introduced by the alkanols in a bilayer.  相似文献   

4.
The interaction of a homologous series of saturated aliphatic n-alkanols (containing 1–13 carbon atoms) with dipalmitoylphosphatidylcholine was studied by a semi-empirical conformational analysis. The minimal conformational energy of the isolated molecule at the hydrocarbon-water interface was calculated as the sum of the contributions resulting from the Van der Waals, torsional, electrostatic and transfer energies. From the conformers of minimal energies were calculated the hydrophilic-hydrophobic balance, the distance between hydrophilic and hydrophobic centres and the energies of interaction between homologous alkanols and between alkanols and lipids. Using these parameters, different modes of conformation, orientation and interaction of n-alkanols and dipalmitoylphosphatidylcholine were described. For methanol, ethanol and n-propanol, the gauche conformers were the most probable interacting only with the lipid polar heads. Only all-trans conformers were obtained for alkanols with longer acyl chains. n-Butanol to n-octanol form clusters in the lipid matrix. Longer n-alkanols are randomly distributed in the lipid layer. However, due to the increase in hydrocarbon chain-length, n-nonanol and higher alkanols have an interaction energy identical or superior to that found in a pure lipid monolayer, leading to a more ordered alkanol-lipid organization.  相似文献   

5.
A series of n-alkanols and phenyl-substituted n-alkanols (Φ-alkanols) of increasing chain length and phenol were characterized for their ability to block action potentials (APs) in frog sciatic nerves. APs were recorded using the single sucrose-gap method. The degree of AP attenuation when the nerve was exposed to different concentrations of an alcohol was used to construct dose-response curves. The reciprocals of the half-blocking doses (ED50s) were used to obtain a measure of the potency of the alcohols. For n-alkanols and Φ-alkanols, increasing the chain length by the addition of a methylene group increased the potency on average by 3.1 for both groups of alkanols. The addition of a phenyl group caused a potency increase that ranged between the values of 77 and 122. The ED50 for both groups of alkanols could not be solely predicted by the log octanol-water partition coefficient (K OW ). Using linear solvation energy relations (LSER), the log ED50 could be described as a linear combination of the intrinsic (van der Waals) molar volume (V I ), polarity (P), and hydrogen bond acceptor basicity (β) and donor acidity (α). Size alone could not predict the ED50 for both n-alkanols and Φ-alkanols. The results are consistent with the hypothesis that alkanols bind to and interact with Na channels to cause AP block. Phenyl group addition to an alkanol markedly increases the molecule's potency. Received: 11 August 2000/Revised: 21 December 2000  相似文献   

6.
Calatayud  A.  Iglesias  D.J.  Talón  M.  Barreno  E. 《Photosynthetica》2004,42(1):23-29
Spinach (Spinacia oleracea L. cv. Clermont) leaves grown in open-top chambers and exposed to three different concentrations of ozone were measured for gas exchange, chlorophyll a fluorescence, antioxidant systems, and lipid peroxidation at the end of growing season. High O3 concentration reduced Fv/Fm, indicating that the efficiency in the energy conversion of photosystem 2 (PS2) was altered. The rate of non-cyclic electron transport rate and the capacity to reduce the quinone pool were also affected. The development of non-photochemical quenching was not high enough to decrease the photon excess in the PS2. The limitation of photosynthetic activity was probably correlated with stomata closure and with an increase in intercellular CO2 concentration. Under oxidative stress, superoxide dismutase (SOD) activity was stimulated in parallel with lipid peroxidation. We did not find any differences in the ascorbate (AsA) pool and ascorbate peroxidase (APX) or glutathione reductase (GR) activities between air qualities. Small, but similar responses were observed in spinach leaves exposed to ambient ozone concentration.  相似文献   

7.
The regulation of photosystem II (PSII) by light-, CO2-, and O2-dependent changes in the capacity for carbon metabolism was studied. Estimates of the rate of electron transport through PSII were made from gas-exchange data and from measurements of chlorophyll fluorescence. At subsaturating photon-flux density (PFD), the rate of electron transport was independent of O2 and CO2. Feedback on electron transport was observed under two conditions. At saturating PFD and low partial pressure of CO2, p(CO2), the rate of electron transport increased with p(CO2). However, at high p(CO2), switching from normal to low p(O2) did not affect the net rate of photosynthetic CO2 assimilation but the rate of electron-transport decreased by an amount related to the change in the rate of photorespiration. We interpret these effects as 1) regulation of ribulose-1,5-bisphosphatecarboxylase (RuBPCase, EC 4.1.1.39) activity to match the rate of electron transport at limiting PFD, 2) regulation of electron-transport rate to match the rate of RuBPCase at low p(CO2), and 3) regulation of the electron-transport rate to match the capacity for starch and sucrose synthesis at high p(CO2) and PFD. These studies provide evidence that PSII is regulated so that the capacity for electron transport is matched to the capacity for other processes required by photosynthesis, such as ribulose-bisphosphate carboxylation and starch and sucrose synthesis. We show that at least two mechanisms contribute to the regulation of PSII activity and that the relative engagement of these mechanisms varies with time following a step change in the capacity for ribulose-bisphosphate carboxylation and starch and sucrose synthesis. Finally, we take advantage of the relatively slow activation of deactivated RuBPCase in vivo to show that the activation level of this enzyme can limit the rate of electron transport as evidenced by increased feedback on PSII following a step change in p(CO2). As RuBPCase as activated, the feedback on PSII declined.Abbreviations and symbols JC electron-transport rate calculated from CO2-assimilation measurements - JF electron-transport rate calculated from fluorescence parameters - PFD photon-flux density - qE energy-dependent quenching - PSII photosystem II - qQ Q-dependent quenching - QY quantum yield - RuBPCase ribulose-1,5-bisphosphate carboxylase (EC 4.1.1.39) C.I.W. publication No. 1015  相似文献   

8.
Abstract— Tryptophan transport across the blood-brain barrier was studied using a single injection dual isotope label technique, in the following three conditions: normal rats, rats with portacaval shunts, and rats with portacaval shunts followed 65 h later by hepatic artery ligation. In both normal rats and those with acute hepatic failure the tryptophan transport system was found to be comprised of two kinetically distinct components. One component was saturable and obeyed Michaelis-Menten kinetics (normal: Vmax= 19.5 nmol.min?1.g?1. Km= 113 μM; hepatic failure: Vmax, = 33.8 nmol.min?1.g?1, Km= 108 μM), and the second was a high capacity system which transported tryptophan in direct proportion to concentration over the range tested (normal: K= 0.026 ml.min?1.g?1; hepatic failure: K= 0.067 ml.min?1.g?1). Since the saturable low capacity component transports several neutral amino acids, and their collective plasma concentration is high in relation to the individual Kms, tryptophan transport by this component is reduced by competitive inhibition under physiological conditions. Thus it was calculated that in normal rats approx 40% of tryptophan influx occurs via the high capacity system. During acute hepatic failure transport via both components was increased substantially, approximately doubling the rate of tryptophan penetration of the blood-brain barrier at all concentrations tested. The contribution by the high capacity component became even more significant than in normal rats, accounting for about 75% of all tryptophan passage from plasma to brain. Brain tryptophan content was 29.9 nmol/g in normal rats and rose to 45.2 nmol/g in rats with portacaval shunts and 50.5 nmol/g in those with acute hepatic failure, correlating with the increased rate of tryptophan transport. In a previous study we found that plasma competing amino acids were greatly increased during acute hepatic failure. Calculations predict that these increased concentrations would cause a reduction in tryptophan transport by the low capacity system. However, because of the increase in the rate of transport by the high capacity component, net tryptophan entry across the blood-brain barrier was actually increased. This increased rate of transport clearly contributes to the increased content of brain tryptophan found during hepatic failure.  相似文献   

9.
In glucose-grown cells of Saccharomyces cerevisiae IGC 4072, acetic acid enters only by simple diffusion of the undissociated acid. In these cells, ethanol and other alkanols enhanced the passive influx of labelled acetic acid. The influx of the acid followed first-order kinetics with a rate constant that increased exponentially with the alcohol concentration, and an exponential enhancement constant for each alkanol was estimated. The intracellular concentration of labelled acetic acid was also enhanced by alkanols, and the effect increased exponentially with alcohol concentration. Acetic acid is transported across the plasma membrane of acetic acid-, lactic acid-, and ethanol-grown cells by acetate-proton symports. We found that in these cells ethanol and butanol inhibited the transport of labelled acetic acid in a noncompetitive way; the maximum transport velocity decreased with alcohol concentration, while the affinity of the system for acetate was not significantly affected by the alcohol. Semilog plots of Vmax versus alcohol concentration yielded straight lines with negative slopes from which estimates of the inhibition constant for each alkanol could be obtained. The intracellular concentration of labelled acid was significantly reduced in the presence of ethanol or butanol, and the effect increased with the alcohol concentration. We postulate that the absence of an operational carrier for acetate in glucose-grown cells of S. cerevisiae, combined with the relatively high permeability of the plasma membrane for the undissociated acid and the inability of the organism to metabolize acetic acid, could be one of the reasons why this species exhibits low tolerance to acidic environments containing ethanol.  相似文献   

10.
The achiral hydrocarbon tetraphenylethylene crystallizes in enantiomorphous forms (chiral space group: P21) to afford right- and left-handed hemihedral crystals, which can be recognized by solid-state circular dichroism spectroscopic analysis. Chiral organic crystals of tetraphenylethylene mediated enantioselective addition of diisopropylzinc to pyrimidine-5-carbaldehyde to give, in conjunction with asymmetric autocatalysis with amplification of chirality, almost enantiomerically pure (S)- and (R)-5-pyrimidyl alkanols whose absolute configurations were controlled efficiently by the crystalline chirality of the tetraphenylethylene substrate. Tetrakis(p-chlorophenyl)ethylene and tetrakis(p-bromophenyl)ethylene also show chirality in the crystalline state, which can also act as a chiral substrate and induce enantioselectivity of diisopropylzinc addition to pyrimidine-5-carbaldehyde in asymmetric autocatalysis to give enantiomerically enriched 5-pyrimidyl alkanols with the absolute configuration correlated with that of the chiral crystals. Highly enantioselective synthesis has been achieved using chiral crystals composed of achiral hydrocarbons, tetraphenylethylenes, as chiral inducers. This chemical system enables significant amplification of the amount of chirality using spontaneously formed chiral crystals of achiral organic compounds as the seed for the chirality of asymmetric autocatalysis.  相似文献   

11.
Ethanol, isopropanol, propanol and butanol enhanced the passive influx of protons into deenergized cells of Saccharomyces cerevisiae. The influx followed first-order kinetics with a rate constant that increased exponentially with the alkanol concentration. The exponential enhancement constants increased with the lipid solubility of the alkanols, which indicated hydrophobic membrane regions as the target sites. While the enhancement constants were independent of pH over the range tested (3.3–5.0), the rate constants decreased linearly with increasing extracellular proton concentration, indicating the presence of an additional surface barrier against proton penetration, the effectiveness of which increased with protonation. The alkanols affected the acidification curves of energized yeast suspensions in such a way that the final pH values were linear functions of the alkanol concentrations. These results were consistent with a balance between active and passive proton movements at the final pH, the exponential enhancement constants calculated from the slopes being nearly identical with those obtained with deenergized cells. It was concluded that passive proton influx contributes to the kinetics of acidification in S. cerevisiae and that uncoupling contributes to the overall kinetics of alkanol-inhibited secondary active transport across the yeast plasma membrane.  相似文献   

12.
The roles of photorespiration and the Mehlerperoxidase pathway in sustaining electron transport and protection from photoinhibition were studied in outer canopy leaves of two species of tropical trees: the drought-deciduous Pseudobombax septenatum (Jacq.) Dug. and the evergreen Ficus insipida Willd. Ficus had a higher photosynthetic capacity than Pseudobombax and also a greater capacity for light-dependent electron transport under photorespiratory conditions (in the absence of CO2). As a consequence, in the absence of CO2, Ficus was able to maintain a largely oxidized electron-transport chain at higher photon flux densities than Pseudobombax. Under the same light conditions, photoinhibition (reduction in Fv/Fm) was always greater in Pseudobombax than Ficus, was increased when leaves were exposed to 2% O2 in nitrogen compared to 21% O2 in CO2-free air, but was not increased by the absence of CO2. Rates of electron transport due to the Mehler-peroxidase pathway (assessed in 2% O2 in nitrogen) ranged between 16–40 mol · m–2·s–1 in both species. As the dry season approached and Pseudobombax neared leaf senescence there was a decline in the capacity for photorespiratory flux to maintain electron transport in Pseudobombax, but not in Ficus. Ratios of light-dependent electron transport to net CO2 fixation for Pseudobombax, Ficus and two other species in the field, Luehea seemannii Tr. & Planch, and Didymopanax morototoni (Aubl.) Dec. & Planch., ranged from 6.2 (Ficus) to 16.7 (Pseudobombax). High in-situ rates of photorespiration combined with the decreased capacity of Pseudobombax for photorespiratory flux as the dry season approached indicates a decreased capacity to protect against photooxidative damage. This may contribute to the promotion of leaf senescence in Pseudobombax during the transition from wet to dry season.Abbreviations Fv/Fm ratio of variable to maximum chlorophyll a fluorescence - NPQ nonphotochemical fluorescence quenching - PFD photon flux density - QA primary electron acceptor of PSII This research was supported by a grant from the Mellon Foundation. We thank Monica Mejia and Juan Posada for assisting with the fluorescence measurements and Aurelio Virgo for assisting with the field CO2-exchange measurements.  相似文献   

13.
Structural and functional characteristics of photosystem II (PSII) were examined in leaves of 4-, 7-, and 11-day-old barley seedlings exposed to high temperature (40°C, 3 h) and water deficit (replacement of nutrient medium with 3% PEG 6000 solution, 45 h). In young seedlings, the effective quantum yield of PSII photochemical reactions decreased upon heat treatment but did not change after dehydration. Both stressful factors diminished the Q B-reducing capacity of PSII in 4- and 7-day-old plants. This was caused by the increase in the reduction level of plastoquinone and by the impairment of the Q B-binding site of the D1 protein. The increase in the content of plastoquinol after the heat shock was due to the impaired oxidizing capacity of cytochrome f (Cyt f). The dehydration did not alter the content of functionally active Cyt f but elevated the microviscosity of the lipid bilayer in thylakoid membranes, which presumably impeded the lateral diffusion of plastoquinones and reduced their reoxidation rate. The heating and dehydration of old leaves reduced the amount of -type PSII reaction centers, thereby inhibiting the linear electron transport. It is concluded that PSII responses to heat treatment and water deficit are subject to variations depending on leaf age. Measurements of the redox potentials for plastoquinones, Cyt f, and cytochrome b 559 upon senescence, hyperthermia, and water deficit allowed us to propose that heat and water stresses activate cyclic electron transport around PSII.  相似文献   

14.
The NMR (nuclear magnetic resonance) method of Conlon and Outhred (1972) was used to measure diffusional water permeability of the nodal cells of the green alga Chara gymnophylla. Two local minima at 15 and 30°C of diffusional water permeability (P d ) were observed delimiting a region of low activation energy (E a around 20 kJ/mol) indicative of an optimal temperature region for membrane transport processes. Above and below this region water transport was of a different type with high E a (about 70 kJ/mol). The triphasic temperature dependence of the water transport suggested a channel-mediated transport at 15–30°C and lipid matrix-mediated transport beyond this region. The K+ channel inhibitor, tetraethylammonium as well as the Cl channel inhibitor, ethacrynic acid, diminished P d in the intermediate temperature region by 54 and 40%, respectively. The sulfhydryl agent p-(chloromercuri-benzensulfonate) the water transport inhibitor in erythrocytes also known to affect K+ transport in Chara, only increased P d below 15°C. In high external potassium (`K-state') water transport minima were pronounced. The role of K+ channels as sensors of the optimal temperature limits was further emphasized by showing a similar triphasic temperature dependence of the conductance of a single K+ channel also known to cotransport water, which originated from cytoplasmic droplets (putatively tonoplast) of C. gymnophylla. The minimum of K+ single channel conductance at around 15°C, unlike the one at 30°C, was sensitive to changes of growth temperature underlining membrane lipid involvement. The additional role of intracellular (membrane?) water in the generation of discontinuities in the above thermal functions was suggested by an Arrhenius plot of the cellular water relaxation rate which showed breaks at 13 and 29°C. Received: 12 August 1998/Revised: 13 November 1998  相似文献   

15.
We investigated how water transport capacity, wood density and wood anatomy were related to leaf photosynthetic traits in two lowland forests in Panama. Leaf-specific hydraulic conductivity (kL) of upper branches was positively correlated with maximum rates of net CO2 assimilation per unit leaf area (Aarea) and stomatal conductance (gs) across 20 species of canopy trees. Maximum kL showed stronger correlation with Aarea than initial kL suggesting that allocation to photosynthetic potential is proportional to maximum water transport capacity. Terminal branch kL was negatively correlated with Aarea/gs and positively correlated with photosynthesis per unit N, indicating a trade-off of efficient use of water against efficient use of N in photosynthesis as water transport efficiency varied. Specific hydraulic conductivity calculated from xylem anatomical characteristics (ktheoretical) was positively related to Aarea and kL, consistent with relationships among physiological measurements. Branch wood density was negatively correlated with wood water storage at saturation, kL, Aarea, net CO2 assimilation per unit leaf mass (Amass), and minimum leaf water potential measured on covered leaves, suggesting that wood density constrains physiological function to specific operating ranges. Kinetic and static indices of branch water transport capacity thus exhibit considerable co-ordination with allocation to potential carbon gain. Our results indicate that understanding tree hydraulic architecture provides added insights to comparisons of leaf level measurements among species, and links photosynthetic allocation patterns with branch hydraulic processes.  相似文献   

16.
Summary It has previously been shown by Macey and Farmer (Biochim. Biophys. Acta 211:104–106, 1970) that phloretin inhibits urea transport across the human red cell membrane yet has no effect on water transport. Jennings and Solomon (J. Gen. Physiol. 67:381–397, 1976) have shown that there are separate lipid and protein binding sites for phloretin on the red cell membrane. We have now found that urea transport is inhibited by phloretin binding to the lipids with aK 1 of 25±8 m in reason-able agreement with theK D of 54±5 m for lipid binding. These experiments show that lipid/protein interactions can alter the conformational state of the urea transport protein. Phloretin binding to the protein site also modulates red cell urea transport, but the modulation is opposed by the specific stilbene anion transport inhibitor, DIDS (4,4-diisothiocyano-2,2-stilbene disulfonate), suggesting a linkage between the urea transport protein and band 3. Neither the lipid nor the protein phloretin binding site has any significant effect on water transport. Water transport is, however, inhibited by up to 30% in a pH-dependent manner by DIDS binding, which suggests that the DIDS/band 3 complex can modulate water transport.  相似文献   

17.
Abstract The kinetics of a range of chlorophyll fluorescence parameters, non-cyclic electron transport and the capacity of the thylakoids to bind Atrazine were examined during photoinhibition treatment of intact pea chloroplasts. Parameters of fluorescence induction of chloroplasts in the presence and absence of 3-(3,4-dichlorophenyl)-1,1-dimethyl urea at 20 °C and at 77 K were determined. The contributions of photochemical and non-photochemical quenching processes to the loss of fluorescence during photoinhibitory treatment were assessed. Two distinct phases of photoinhibitory damage were observed. During the initial 5 min period of exposure to light the minimal fluorescence level (Fo) increased, whilst the maximal fluorescence level (FP) decreased, both coupled and uncoupled non-cyclic electron transport to methyl viologen decreased and the ability to bind Atrazine to the thylakoids decreased. Fluorescence analyses demonstrated that during this period thylakoids were becoming increasingly less efficient at generating and maintaining a transmembrane proton electrochemical gradient. Photoinhibitory damage that occurred at later times between 5 and 20 min was of a very different nature. Both Fo and FP declined, a loss of coupled and uncoupled non-cyclic electron transport was observed together with a loss of the capacity to photo-oxidize water. However, no further loss of Atrazine-binding was associated with such changes. A consistent decrease in the quantum yield of non-cyclic electron transport was also observed throughout photoinhibition treatment. The possibility of two distinct mechanisms of photoinhibitory damage to the photosynthetic apparatus is discussed.  相似文献   

18.
To study the role the phenyl group plays in producing local anesthetic block, a sequence of n-alkanols and phenyl-substituted alkanols (Φ-alkanols) were characterized in their ability to block Na channels. The sequence of n-alkanols studied possess 3–5 carbons (propanol-pentanol). The action of phenol and 3-Φ-alkanols (benzyl alcohol, phenethyl alcohol, 3-phenyl-1-propanol) were also studied. Na currents (I Na ) were recorded from single frog skeletal muscle fibers using the Vaseline-gap voltage clamp technique. I Na s were recorded prior to, during, and following the removal of the solutes in Ringer's solution. All alkanols and phenol acted to block I Na in a dose-dependent manner. Effective doses to produce half block (ED50) of I Na or Na conductance (G Na ) were obtained from dose-response relations for all solutes used. The block of G Na depended on voltage, and could be separated into voltage-dependent and -independent components. Each solute acted to shift G Na -V relations in a depolarized direction and reduce the maximum G Na and slope of the relation. All solutes acted to speed up I Na kinetics and cause hyperpolarizing shifts in steady-state inactivation. The magnitude of the kinetic changes increased with dose. Size was an important variable in determining the magnitude of the changes in I Na ; however, size alone was not sufficient to predict the changes in I Na . ED50s for G Na and AP block could be predicted as a function of intrinsic molar volume, hydrogen bond acceptor basicity (β) and donor acidity (α), and polarity (P) of the solutes. The equivalency of ED50 predictions for AP and G Na block can be explained by the fact that AP block arises from channel block and solute-induced changes in I Na kinetics. Φ-alkanols were more effective at blocking and inactivating Na channels than their unsubstituted counterparts. Phenyl-substituted alkanols are more likely to interact with the channel than their unsubstituted counterparts. Received: 11 August 2000/Revised: 21 December 2000  相似文献   

19.
The temperature sensitive release and uptake of ATP through theAspergillus niger G3Br membrane vesicles followed saturation kinetics. Both the processes which occurred in the absence of mycobacillin were greatly enhanced by its presence. Liposomes prepared with antifilipin sterol and lipid showed the release and uptake of ATP in the presence of filipin, but no such uptake and release was seen with antimycobacillin sterol and lipid in the presence of mycobacillin. However the liposomes supplemented withAspergillus niger membranes protein (s) showed the release and uptake of ATP, implicating membrane protein as a carrier in the transport process.  相似文献   

20.
Corynebacterium equi IFO 3730 was found to oxidize a wide variety of sec-alcohols, including alkanols, substituted alkanols, alkenols and cyclic alcohols, in moderate to high yields. Among them, the sec-alcohols which have longer carbon chains were oxidized more smoothly than those with smaller numbers of carbon. Although both enantiomers of unsymmetrically disubstituted carbinols were oxidized, the S form of 2-dodecanol was converted to the corresponding ketone a little faster than the other enantiomer.  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