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.
Summary Volume-dependent changes in light scatter have been shown to be a linear function of the osmotic gradient imposed upon gastric vesicles purified from hog gastric mucosa. Observation of the light scattered 90° to incident, using the Durrum stop flow system D-110, indicates that the vesicles exposed to hypertonic medium undergo rapid shrinkage due to water loss from the vesicle interior. The rate constant for this water movement is 1.1±0.09 sec–1 (n=10) and is linearly dependent on temperature between 16 and 36°C. The activation energy of 13.93±0.60 kcal mole–1 (n=3), calculated from an Arrhenius plot, is inconsistent with water movement facilitated by a large-pore aqueous channel. A slower reswell phase, dependent on solute entry into the intravesicular space, follows the water-dependent shrink phase. KCl entry, studied because of the intravesicular requirement for active K+/H+ transport, exhibits two entry stages. The faster, described by a single exponential imposed upon a constantly sloping background, has a rate constant of 7.75±0.48×10–3 sec–1 (n=15). The slower phase, which typically accounts for 90% of the reswell process, demonstrates a rate constant of 1.94±0.23×10–4 sec–1 (n=15). In the presence of valinomycin or nigericin, two fast rate constants and one slow rate constant of swelling are observed. The rate constant of the faster reswell phase is increased from 7.75±0.48×10–3 sec–1 (n=15) to 15.74±3.7×10–3 sec–1 (n=5) and 17.23±3.4×10–3 (n=3) by the addition of nigericin (1 g ml–1) and valinomycin (4.5 m), respectively. The second part of the faster reswell phase is approximately that seen in the control population. Transport-dependent volume changes of significant magnitude can be demonstrated following the addition of ATP to vesicles equilibrated with 150mm KCl. The volume change is a function of HCl leak rate and is abolished by ionophores which eliminate the transport-dependent pH gradient. So 4 –- substitution, which eliminates the overshoot phenomena observed in KCl medium, also eliminates the shrinkage resulting from ATP addition.  相似文献   

2.
Summary Permeabilities of ammonia (NH3), methylamine (CH3NH2) and ethylamine (CH3CH2NH2) in the cyanobacterium (cyanophyte)Synechococcus R-2 (Anacystis nidulans) have been measured. Based on net uptake rates of DCMU (dichlorophenyldimethylurea) treated cells, the permeability of ammonia was 6.44±1.22 m sec–1 (n=13). The permeabilities of methylamine and ethylamine, based on steady-state14C labeling were more than ten times that of ammonia (P methylamine=84.6±9.47 m sec–1 (76),P ethylamine=109±11 m sec–1 (55)). The apparent permeabilities based on net uptake rates of methylamine and ethylamine uptake were significantly lower, but this effect was partially reversible by ammonia, suggesting that net amine fluxes are rate limited by proton fluxes to an upper limit of about 700 nmol m–2 sec–1. Increasing concentrations of amines in alkaline conditions partially dissipated the pH gradient across the cell membrane, and this property could be used to calculate the relative permeabilities of different amines. The ratio of ethylamine to methylamine permeabilities was not significantly different from that calculated from the direct measurements of permeabilities; ammonia was much less effective in dissipating the pH gradient across the cell membrane than methylamine or ethylamine. An apparent permeability of ammonia of 5.7±0.9 m sec–1 could be calculated from the permeability ratio of ammonia to methylamine and the experimentally measured permeability of methylamine. The permeability properties of ammonia and methylamine are very different; this poses problems in the interpretation of experiments where14C-methylamine is used as an ammonia analogue.  相似文献   

3.
Summary The effects of stepwise concentration changes of K+ and HCO 3 in the basolateral solution on the basolateral membrane potential (V bl) of proximal tubule cells of the doubly-perfusedNecturus kidney were examined using conventional microelectrodes. Apparent transference numbers were calculated from changes inV bl after alterations in external K+ concentration from 1.0 to 2.5mm (t K, 1.0–2.5), 2.5 to 10, and in external HCO 3 concentration (at constant pH) from 5 to 10mm (t HCO3, 5–10), 10 to 20, or 10 to 50.t K, 2.5–10 was 0.38±0.02 under control conditions but was sharply reduced to 0.08±0.03 (P>0.001) by 4mm Ba++. This concentration of Ba++ reducedV bl by 9±1 mV (at 2.5 external K+). Perfusion with SITS (5×10–4 m) for 1 hr hyperpolarizedV bl by 10±3 mV and increasedt K, 2.5–10 significantly to 0.52±0.01 (P<0.001). Ba++ application in the presence of SITS depolarizedV bl by 22±3 mV. In control conditionst HCO3, 10–50 was 0.63±0.05 and was increased to 0.89±0.07 (P<0.01) by Ba++ but was decreased to 0.14±0.02 (P<0.001) by SITS. In the absence of apical and basolateral chloride, the response ofV bl to bicarbonate was diminished but still present (t HCO3, 10–20 was 0.35±0.03). Intracellular pH, measured with liquid ion-exchange microelectrodes, increased from 7.42±0.19 to 7.57±0.17 (P<0.02) when basolateral bicarbonate was increased from 10 to 20mm at constant pH. These data show that the effects of bicarbonate onV bl are largely independent of effects on the K+ conductance and that there is a significant current-carrying bicarbonate pathway in the basolateral membrane. Hence, both K+ and HCO 3 gradients are important in the generation ofV bl, and their relative effects vary reciprocally.  相似文献   

4.
Summary Na+–H+ exchange activity in renal brush border membrane vesicles isolated from hyperthyroid rats was increased. When examined as a function of [Na+], treatment altered the initial rate of Na+ uptake by increasingV m (hyperthyroid, 18.9±1.1 nmol Na+ · mg–1 · 2 sec–1; normal, 8.9±0.3 nmol Na+ · mg–1 · 2 sec–1), and not the apparent affinityK Na + (hyperthyroid, 7.3±1.7mm; normal, 6.5±0.9mm). When examined as a function of [H+] and at a subsaturating [Na+] (1mm), hyperthyroidism resulted in the proportional increase in Na+ uptake at every intravesicular pH measured. A positive cooperative effect on Na+ uptake was found with increased intravesicular acidity in vesicles from both normal and hyperthyroid rats. When the data were analyzed by the Hill equation, it was found that hyperthyroidism did not change then (hyperthyroid, 1.2±0.06; normal, 1.2±0.07) or the [H+]0.5 (hyperthyroid, 0.39±0.08 m; normal, 0.44±0.07 m) but increased the apparentV m (hyperthyroid, 1.68±0.14 nmol Na+ · mg–1 · 2 sec–1; normal 0.96±0.10 nmol Na+ · mg–1 · 2 sec–1). The uptake of Na+ in exchange for H+ in membrane vesicles from normal and hyperthyroid animals was not influenced by membrane potential. H+ translocation or debinding was rate limiting for Na+–H+ exchange since Na+–Na+ exchange activity was greater than Na+–H+ exchange activity. Hyperthyroidism caused a proportional increase and hypothyroidism caused a proportional decrease in Na+–Na+ and Na+–H+ exchange. We conclude that hyperthyroidism leads to either an increase in the number of functional exchangers in the membrane or exactly proportional increases in the rate-limiting steps for Na+–Na+ and Na+–H+ exchange activity.  相似文献   

5.
Summary Cl transport in apical membrane vesicles derived from bovine tracheal epithelial cells was studied using the Cl-sensitive fluorescent indicator 6-methoxy-N-(3-sulfopropyl) quinolinium. With an inwardly directed 50 mM Cl gradient at 23°C, the initial rate of Cl entry (J Cl) was increased significantly from 0.32±0.12 nmol · sec–1 · mg protein–1 (mean±sem) to 0.50±0.07 nmol · sec–1 · mg protein–1 when membrane potential was changed from 0 to +60 mV with K/valinomycin. At 37°C, with membrane potential clamped at 0 mV, there was a 34±7% (n=5) decrease inJ Cl from a control value of 0.37±0.03 nmol · sec–1 · mg protein–1 upon addition of 0.2mm diphenylamine-2-carboxylate. The following did not alterJ Cl significantly (J Cl values gives as percent change from control): 50mm cis Na (–1±5%), 0.1mm furosemide (–3±4%), 0.1mm furosemide in the presence of 50mm cis Na (–5±2%), 0.1mm H2DIDS (–18±9%), a 1.5 pH unit inwardly directed H gradient (–7±7%), and 0.1mm H2DIDS in the presence of a 1.5 unit pH gradient (4±18%). With inward 50mm anion gradients, the initial rates of Br and I entry (J Br andJ 1, respectively) were not significantly different fromJ Cl.J Cl was a saturable function of Cl concentration with apparentK d of 24mm and apparentV max of 0.54 nmol · sec–1 · mg protein–1. Measurement of the temperature dependence ofJ Cl yielded an activation energy of 5.0 kcal/mol (16–37°C). These results demonstrate that Cl transport in tracheal apical membrane vesicles is voltage-dependent and inhibited by diphenylamine-2-carboxylate. There is no significant contribution from the Na/K/2Cl, Na/Cl, or Cl/OH(H) transporters. The conductive pathway does not discriminate between Cl, Br, and I and is saturable. The low activation energy supports a pore-type mechanism for the conductance.  相似文献   

6.
Summary The transepithelial water permeability in frog urinary bladder is believed to be essentially dependent on the ADH-regulated apical water permeability. To get a better understanding of the transmural water movement, the diffusional water permeability (P d) of the basolateral membrane of urinary bladder was studied. Access to this post-luminal barrier was made possible by perforating the apical membrane with amphotericin B. The addition of this antibiotic increasedP d from 1.12±0.10×10–4 cm/sec (n=7) to 4.08±0.33×10–4 cm/sec (n=7). The effect of mercuric sulfhydryl reagents, which are commonly used to characterize water channels, was tested on amphotericin B-treated bladders. HgCl2 (10–3 m) decreasedP d by 52% andpara-chloromercuribenzoic acid (pCMB) (1.4×10–4 m) by 34%. The activation energy for the diffusional water transport was found to increase from 4.52±0.23 kcal/mol (n=3), in the control situation, to 9.99±0.91 kcal/mol (n=4) in the presence of 1.4×10–4 m pCMB. Our second approach was to measure the kinetics of water efflux, by stop-flow light scattering, on isolated epithelial cells from urinary bladders.pCMB (0.5 or 1.4×10–4 m) was found to inhibit water exit by 91±2%. These data strongly support the existence of proteins responsible for water transport across the basolateral membrane, which are permanently present.  相似文献   

7.
Summary Na+ and sugar permeabilities of egg lecithin bilayers were measured using curved bilayers and planar bilayers as represented by single-bilayer vesicles and black lipid films, respectively. The Na+ permeability coefficient measured with single-bilayer vesicles at 25°C is (2.1±0.6)×10–13 cm sec–1. Because of technical difficulties it has been impossible to measure ionic permeabilities of values lower than about 10–10 cm sec–1 in planar (black) lipid bilayers using tracer methods. Thed-glucose andd-fructose permeabilities were measured with both curved and planar bilayers. The permeability coefficients measured with vesicles at 25°C are (0.3±0.2)×10–10 cm sec–1 for glucose and (4±1)×10–10 cm sec–1 ford-fructose; these are in reasonable agreement with the corresponding values obtained for planar (black) lipid bilayers which are (1.1±0.3)×10–10 cm sec–1 ford-glucose and (9.3±0.3)×10–10 cm sec–1 ford-fructose, respectively.This paper is dedicated to the memory of Walther Wilbrandt,cuius nomini nullum par elogium.  相似文献   

8.
The interaction of three forms of bovine angiotensin-converting enzyme (ACE) with the competitive peptide inhibitor lisinopril with a fluorescent label was studied by the fluorescence polarization technique. The dissociation constants K d of the enzyme-inhibitor complexes in 50 mM Hepes-buffer, pH 7.5, containing 150 mM NaCl and 1 M ZnCl2 at 37°C were (2.3 ± 0.4)·10–8, (2.1 ± 0.3)·10–8, and (2.1 ± 0.2)·10–8 M for two-domain somatic ACE, single-domain testicular ACE, and for the N-domain of the enzyme, respectively. The interaction of the enzyme with the inhibitor strongly depended on the presence of chloride in the medium, and the apparent dissociation constant of the ACE-chloride complex was (1.3 ± 0.2)·10–3 M for the somatic enzyme. The dissociation kinetics of the complex of the inhibitor with somatic ACE did not fit the kinetics of a first-order reaction, but it was approximated by a model of simultaneous dissociation of two complexes with the dissociation rate constants (0.13 ± 0.01) sec–1 and (0.026 ± 0.001) sec–1 that were present at approximately equal initial concentrations. The dissociation kinetics of the single-domain ACE complexes with the inhibitor were apparently first-order, and the dissociation rate constants were similar: (0.055 ± 0.001) and (0.041 ± 0.001) sec–1 for the N-domain and for testicular ACE, respectively.  相似文献   

9.
The mechanism of nitrate transport across the tonoplast of barley root cells   总被引:14,自引:0,他引:14  
Nitrate-selective microelectrodes were used to measure not only nitrate activity in the cytoplasm and vacuole of barley (Hordeum vulgare L.) root cells, but also the tonoplast electrical membrane potential. For epidermal cells, the mean cytoplasmic and vacuolar pNO3 (-log10 [NO3]) values were 2.3±0.04 (n=19) and 1.41±0.03 (n=35), respectively, while for cortical cells, the mean cytoplasmic and vacuolar nitrate values were 2.58±0.18 (n=4) and 1.17±0.06 (n=13), respectively. These results indicate that the accumulation of nitrate in the vacuole must be an active process. Proton-selective microelectrodes were used to measure the proton gradient across the tonoplast to assess the possibility that nitrate transport into the vacuole is mediated by an H+/NO 3 antiport mechanism. For epidermal cells, the mean cytoplasmic and vacuolar pH values were 7.12±0.06 (n=10) and 4.93±0.11 (n=22), respectively, while for cortical cells, the mean cytoplasmic and vacuolar pH values were 7.24±0.07 (n=3) and 5.09±0.17 (n=7), respectively. Calculations of the energetics for this mechanism indicate that the observed gradient of nitrate across the tonoplast of both epidermal and cortical cells could be achieved by an H+/NO 3 antiport with a 11 stoichiometry.Abbreviations and Symbols G/F free-energy change for H+/NO 3 antiport - F Faraday constant - pHc cytoplasmic pH - pHv vacuolar pH - p[NO3]c log10 (cytoplasmic [NO 3 ]) - P[NO3]v -log10 (vacuolar [NO3]) We wish to thank Dr. K. Moore for assistance with statistical analysis.  相似文献   

10.
Summary Transport of protons and halide ions through planar lipid bilayers made from egg lecithin and a long-chain secondary amine (n-lauryl [trialkylmethyl] amine) inn-decane was studied. Net proton fluxes were measured with a pH electrode, and halide fluxes were measured with82Br and36Cl. In membranes containing the secondary amine, a large net proton flux was produced either by a Br gradient with symmetrical pH or by a pH gradient with symmetrical Br, but not by a pH gradient in Br-free solutions. This H+ flux was electrically silent (nonconductive), and the H+ permeability coefficient was >10–3 cm sec–1 in 0.1m NaBr. In Br-free solutions, H+ selectivity was observed electrically by measuring conductances and zero-current potentials generated by H+ activity gradients. The permeability coefficient for this ionic (conductive) H+ flux was about 10–5 cm sec–1, several orders of magnitude smaller than the H+ permeability of the electroneutral pathway. Large electroneutral Br exchange fluxes occurred under symmetrical conditions, and the permeability coefficient for Br exchange was about 10–3 cm sec–1 at pH 5. The one-way Br flux was inhibited by substituting SO 4 = for Br on the trans side of the membrane. These results support a titratable carrier model in which the secondary amine exists in three forms (C, CH+ and CHBr). Protons can cross the membrane either as CHBr (nonconductive) or as CH+ (conductive), whereas Br crosses the membrane primarily as CHBr (nonconductive). In addition to these three types of transport, there is also a pH-dependent conductive flux of Br which has a permeability coefficient of about 10–7 cm sec–1 at pH 5. Experiments with lipid monolayers suggest that the pH dependence of this conductive flux is caused by a change in surface potential of about +100 mV between pH 9.5 and 5.0.  相似文献   

11.
Summary Water transport across the mammalian collecting tubule is regulated by vasopressin-dependent water channel insertion into and retrieval from the cell apical membrane. The time course of osmotic water permeability (P f ) following addition and removal of vasopressin (VP) and 8-Br-cAMP was measured continuously by quantitative fluorescence microscopy using an impermeant fluorophore perfused in the lumen. Cortical collecting tubules were subjected to a 120 mOsm bath-to-lumen osmotic gradient at 37°C with 10–15 nl/min lumen perfusion and 10–20 ml/min bath exchange rate. With addition of VP (250 U/ml), there was a 23±3 sec (sem,n=16) lag in whichP f did not change, followed by a rise inP f (initial rate 1.4±0.2×10–4 cm/sec2) to a maximum of 265±10×10–4 cm/sec. With addition of 8-Br-cAMP (0.01–1mm) there was an 11±2 sec lag. For [8-Br-cAMP]=0.01, 0.1 and 1mm, the initial rate ofP f increase following the lag was (units 10–4 cm/sec2): 1.1±0.1, 1.2±0.1 and 1.7±0.3. MaximumP f was (units 10–4 cm/sec): 64±4, 199±9 and 285±11. With removal of VP,P f decreased to baseline (12×10–4 cm/sec) with aT 1/2 of 18 min; removal of 0.1 and 1mm 8-Br-cAMP gaveT 1/2 of 4 and 8.5 min. These results demonstrate (i) a brief lag in theP f response, longer for stimulation by VP than by 8-Br-cAMP, representing the transient build-up of biochemical intermediates proximal to the water channel insertion step, (ii) similar initialdP f /dt (water channel insertion) over a wide range of [8-Br-cAMP] and steady-stateP f values, and (iii) more rapidP f decrease with removal of 8-Br-cAMP than with VP. These pre-steady-state results define the detailed kinetics of the turn-on and turn-off of tubuleP f and provide kinetic evidence that the rate-limiting step for turn-on ofP f is not the step at which VP regulates steady-stateP f . If water channel insertion is assumed to be the rate-limiting step in the turn-on ofP f , these results raise the possibility that water channels must be activated following insertion into the apical membrane.  相似文献   

12.
Summary The kinetics of the separate processes of Fe2(III)-transferrin binding to the transferrin receptor, transferrin-receptor internalization, iron dissociation from transferrin, iron passage through the membrane, and iron mobilization into the cytoplasm were studied by pulse-chase experiments using rabbit reticulocytes and59Fe,125I-labeled rabbit transferrin. The binding of59Fe-transferrin to transferrin receptors was rapid with an apparent rate constant of 2×105 m –1 sec–1. The rate of internalization of59Fe-transferrin was directly measured at 520±100 molecules of Fe2(III)-transferrin internalized/sec/cell with 250±43 sec needed to internalize the entire complement of reticulocyte transferrin receptors. Subsequent to Fe2(III)-transferrin internalization the flux of59Fe was followed through three compartments: internalized transferrin, membrane, and cytosol.A process preceding iron dissociation from transferrin and a reaction involving membrane-associated iron required 17±2 sec and 34±5 sec, respectively. Apparent rate constants of 0.0075±0.002 sec–1 and 0.0343±0.0118 sec–1 were obtained for iron dissociation from transferrin and iron mobilization into the cytosol, respectively. Iron dissociation from transferrin is the rate-limiting step. An apparent rate constant of 0.0112±0.0025 sec–1 was obtained for processes involving iron transport through the membrane although at least two reactions are likely to be involved. Based on mechanistic considerations, iron transport through the membrane may be attributed to an iron reduction step followed by a translocation step. These data indicate that the uptake of iron in reticulocytes is a sequential process, with steps after the internalization of Fe2(III)-transferrin that are distinct from the handling of transferrin.  相似文献   

13.
Summary The shifts of current-voltage characteristics of sodium and calcium inward currents produced by changes in the concentration of divalent cations (Mg2+, Ca2+, Sr2+, Ba2+) and in pH of the extracellular solution have been measured on isolated neurons of the molluscHelix pomatia intracellularly perfused with potassium-free solutions. On the basis of these shifts and using Stern's theory (O. Stern, 1924.Z. Electrochem. 30508–516), the binding constants for the ions to charged groups of the outer side of the somatic membrane and the density of the surface charges produced by these groups have been calculated. For groups located in the vicinity of sodium channels we obtainedK Ca=90±10,K Sr=60±10,K Ba=25±5 andK Mg=16±5m –1 at pH=7.7 and for groups located in the vicinity of calcium channelsK Ca=67±10,K Sr=20±5 andK Ba=19±5m –1 at pH=7.0. The same groups bind H+ ions with apparent pK=6.2±0.2 that corresponds toK H=1.6×106 m –1. The density of fixed charges near the sodium channels is 0.17±0.05 e/nm2 (pH=7.7) and near the calcium channels is 0.23±0.05 electrons/nm2 (pH=7.0). From the comparison of the obtained values with the data about binding constants of the same ions to different negatively charged phospholipids, a suggestion is made that just the phophatidylserine is responsible for the surface potential of the outer side of the somatic membrane. It was also shown that the presence of this potential results in a change in the concentration of carrier ions near the membrane which affects the maximal values of the corresponding transmembrane currents.  相似文献   

14.
Summary Confluent monolayers of the established opossum kidney cell line were exposed to NH4Cl pulses (20 mmol/liter) during continuous intracellular measurements of pH, membrane potential (PD m ) and membrane resistance (R m) in bicarbonate-free Ringer. The removal of extracellular NH4Cl leads to an intracellular acidification from a control value of 7.33±0.08 to 6.47±0.03 (n=7). This inhibits the absolute K conductance (g K+), reflected by a decrease of K+ transference number from 71±3% (n=28) to 26±6% (n=5), a 2.6±0.2-fold rise ofR m, and a depolarization by 24.2±1.5 mV (n=52). In contrast, intracellular acidification during a block ofg K+ by 3 mmol/liter BaCl2 enhances the total membrane conductance, being shown byR m decrease to 68±7% of control and cell membrane depolarization by 9.8±2.8 mV (n=17). Conversely, intracellular alkalinization under barium elevatesR m and hyperpolarizes PD m . The replacement of extracellular sodium by choline in the presence of BaCl2 significantly hyperpolarizes PD m and increasesR m, indicating the presence of a sodium conductance. This conductance is not inhibited by 10–4 mol/liter amiloride (n=7). Patch-clamp studies at the apical membrane (excised inside-out configuration) revealed two Na+-conductive channels with 18.8±1.4 pS (n=10) and 146 pS single-channel conductance. Both channels are inwardly rectifying and highly selective towards Cl. The low-conductive channel is 4.8 times more permeable for Na+ than for K+. Its open probability rises at depolarizing potentials and is dependent on the pH of the membrane inside (higher at pH 6.5 than at pH 7.8).  相似文献   

15.
Summary Transport by an epithelium, possessing an accumulating, saturable transport system in the apical membrane as well as a finite Fick permeability to the transported solute, was considered in the steady state in the case of zerocis concentration, and in the presence of a peripheral diffusion resistance in a layer apposing thecis face of the tissue (unstirred solution or structural coating). Under suitable conditions, the combination of peripheral diffusion resistance and accumulating epithelial transport may lead to recycling of solute at thecis face of the epithelium. This causes a decrease of the effective permeability to diffusionaltrans-cis flow across the tissue. The phenomenon is discussed in terms of epidermald-glucose transport by the integument of aquatic animals with a collagenous cuticle, such as the seawater-acclimated polychaete wormNereis diversicolor. The recycling phenomenon may be of significance to other epithelia with the function of maintaining large concentration gradients of permeating substances.List of Symbols and Fixed Parameter Values C m Bulk medium solute concentration,cis face of epidermisC m=0 mol cm–3 - C i Concentration of solute at interface between cuticle and unstirred medium (mol cm–3) - C s Concentration of solute atcis face of apical epidermal membrane (mol cm–3) - C e Concentration of solute in extracellular fluid,trans-side of epidermisC e=1.0×10–6 mol cm–3 - D m Diffusion coefficient of solute in outside mediumD m=6.7×10–6 cm2 sec–1 - D c Diffusion coefficient of solute in cuticleD c=7.4×10–9 cm2 sec–1 - m Operative thickness of unstirred medium layer - c Thickness of cuticle - J Steady-state net flux of solute through cuticle or unstirred layer (flux is positive indirectioncis-trans) (mol cm–2 sec–1) - J i max Maximal influx through saturable transport system in apical membraneJ i max =2.0×10–12 mol cm–2 sec–1 - K t Transport constant, saturable systemK t=1.0×10–7 mol cm–3 - P Epithelial permeability (cm sec–1)  相似文献   

16.
Summary Methods have been used for monitoring either volume flows or pressure changes, simultaneously with membrane potentials, in giant algal cells ofChara australis during an action potential. The volume flows were measured from the movement of a mercury bead in a capillary tube recorded by a photo-transducer. The pressure changes were measured by monitoring the deflection of a thin wedge, resting transversely across a cell, and using the same photo-transducer, the deflection of the wedge being directly related to the cell's turgor pressure. The average maximum rate of volume flow per unit area during an action potential was 0.88±0.11 nliter·sec–1·cm–2 in the direction of an outflow from the cell (total volume outflow being about 3 nliter·cm–2 per action potential). Similarly, the maximum rate of change of pressure was 19.6±3.8×10–3 atm·sec–1 (peak change being 19.3±2.9×10–3 atm equivalent to 14.7±2.2 mm Hg). The volume flow and pressure changes followed the vacuolar potential quite closely, the peak rate of volume flow lagging behind the peak of the action potential by 0.17±0.08 sec and the peak rate of pressure change leading it by 0.09±0.07 sec.  相似文献   

17.
Two trypsin inhibitors, LA-1 and LA-2, have been isolated from ridged gourd (Luffa acutangula Linn.) seeds and purified to homogeneity by gel filtration followed by ion-exchange chromatography. The isoelectric point is atpH 4.55 for LA-1 and atpH 5.85 for LA-2. The Stokes radius of each inhibitor is 11.4 å. The fluorescence emission spectrum of each inhibitor is similar to that of the free tyrosine. The biomolecular rate constant of acrylamide quenching is 1.0×109 M–1 sec–1 for LA-1 and 0.8 × 109 M–1 sec–1 for LA-2 and that of K2HPO4 quenching is 1.6×1011 M–1 sec–1 for LA-1 and 1.2×1011M–1 sec–1 for LA-2. Analysis of the circular dichroic spectra yields 40%-helix and 60%-turn for La-1 and 45%-helix and 55%-turn for LA-2. Inhibitors LA-1 and LA-2 consist of 28 and 29 amino acid residues, respectively. They lack threonine, alanine, valine, and tryptophan. Both inhibitors strongly inhibit trypsin by forming enzymeinhibitor complexes at a molar ratio of unity. A chemical modification study suggests the involvement of arginine of LA-1 and lysine of LA-2 in their reactive sites. The inhibitors are very similar in their amino acid sequences, and show sequence homology with other squash family inhibitors.  相似文献   

18.
Summary The potential dependence of unidirectional36Cl fluxes through toad skin revealed activation of a conductive pathway in the physiological region of transepithelial potentials. Activation of the conductance was dependent on the presence of Cl or Br in the external bathing solution, but was independent of whether the external bath was NaCl-Ringer's, NaCl-Ringer's with amiloride, KCl-Ringer's or choline Cl-Ringer's To partition the routes of the conductive Cl ion flow, we measured in the isolated epithelium with double-barrelled microelectrodes apical membrane potentialV a , and intracellular Cl activity,a Cl c , of the principal cells indentified by differential interference contrast microscopy. Under short-circuit conditionsI sc=27.0±2.0 A/cm2, with NaCl-Ringer's bathing both surfaces,V a was –67.9±3.8mV (mean ±se,n=24, six preparations) anda Cl c was 18.0±0.9mM in skins from animals adapted to distilled water. BothV a anda Cl a were found to be positively correlated withI sc (r=0.66 andr=0.70, respectively). In eight epithelia from animals adapted to dry milieu/tap waterV a anda Cl c were measured with KCl Ringer's on the outside during activation and deactivation of the transepithelial Cl conductance (G Cl) by voltage clamping the transepithelial potential (V) at 40 mV (mucosa positive) and –100 mV. AtV=40 mV; i.e. whenG Cl was deactivated,V a was –70.1±5.0 mV (n=15, eight preparations) anda Cl c was 40.0±3.8mm. The fractional apical membrane resistance (fR a) was 0.69±0.03. Clamping toV=–100 mV led to an instantaneous change ofV a to 31.3±5.6 mV (cell interior positive with respect to the mucosal bath), whereas neithera Cl c norfR a changed significantly within a 2 to 5-min period during whichG Cl increased by 1.19±0.10 mS/cm2. WhenV was stepped back to 40 mV,V a instantaneously shifted to –67.8±3.9 mV whilea Cl c andfR a remained constant during deactivation ofG Cl. Similar results were obtained in epithelia impaled from the serosal side. In 12 skins from animals adapted to either tap water or distilled water the density of mitochondria-rich (D MRC) cells was estimated and correlated with the Cl current (I Cl though the fully activated (V=–100mV) Cl conductance). A highly significant correlation was revealed (r=–0.96) with a slope of –2.6 nA/m.r. (mitochondria-rich cell and an I-axis intercept not significantly different from zero. In summary, the voltage-dependent Cl currents were not reflected infR a anda Cl a of the principal cells but showed a correlation with the m.r. cell density. We conclude that the pricipal cells do not contribute significantly to the voltage-dependent Cl conductance.  相似文献   

19.
Erythrocyte membrane potential can be estimated by measuring the transmembrane concentration (activity) distribution of a membrane-permeable ion. We present here the study of difluorophosphate (DFP) as a 19F NMR probe of membrane potential. This bicarbonate and phosphate analogue has a pKa of 3.7±0.2 (SD, n = 4) and therefore exists almost entirely as a monovalent anion at physiological pH. When it is incorporated into red cell suspensions, it gives two well resolved resonances that arise from the intra- and extracellular populations; the intracellular resonance is shifted 130 Hz to higher frequency from that of the extracellular resonance. Hence the transmembrane distribution of DFP is readily assessed from a single 19F NMR spectrum and the membrane potential can be calculated using the Nernst equation. The membrane potential was independent of, DFP concentration in the range 4 to 59 mM, and haematocrit of the cell suspensions of 31.0 to 61.4%. The membrane potential determined by using DFP was 0.94±0.26 of that estimated from the transmembrane pH difference. The distribution ratios of intracellular/extracellular DFP were similar to those of the membrane potential probes, hypophosphite and trifluoroacetate. DFP was found to be transported across the membranes predominantly via the electrically-silent pathway mediated by capnophorin. Using magnetization transfer techniques, the membrane influx permeability-coefficient of cells suspended in physiological medium was determined to be 7.2±2.5 × 10–6 cm s–1 (SD, n=4). Offprint requests to: P. W Kuchel  相似文献   

20.
Summary The effect of chloride on 4,4-dibenzamido-2,2-disulfonic stilbene (DBDS) binding to band 3 in unsealed red cell ghost membranes was studied in buffer [NaCl (0 to 500mm) + Na citrate] at constant ionic strength (160 or 600mm). pH 7.4, 25°C. In the presence of chloride, DBDS binds to a single class of sites on band 3. At 160mm ionic strength, the dissociation constant of DBDS increases linearly with chloride concentration in the range [Cl]=450mm. The observed rate of DBDS binding to ghost membranes, as measured by fluorescence stopped-flow kinetic experiments, increases with chloride concentration at both 160 and 600mm ionic strength. The equilibrium and kinetic results have been incorporated into the following model of the DBDS-band 3 interaction: The equilibrium and rate constants of the model at 600mm ionic strength areK 1=0.67±0.16 m,k 2=1.6±0.7 sec–1,k –2=0.17±0.09 sec–1,K 1=6.3±1.7 m,k 2=9±4 sec–1 andk –2=7±3 sec–1. The apparent dissociation constants of chloride from band 3,K Cl, are 40±4mm (160mm ionic strength) and 11±3mm (600mm ionic strength). Our results indicate that chloride and DBDS have distinct, interacting binding sites on band 3.  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