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 53 毫秒
1.
E. Steudle  J. S. Boyer 《Planta》1985,164(2):189-200
Hydraulic resistances to water flow have been determined in the cortex of hypocotyls of growing seedlings of soybean (Glycine max L. Merr. cv. Wayne). Data at the cell level (hydraulic conductivity, Lp; half-time of water exchange, T 1/2; elastic modulus, ; diffusivity for the cell-to-cell pathway, D c) were obtained by the pressure probe, diffusivities for the tissue (D t) by sorption experiments and the hydraulic conductivity of the entire cortex (Lpr) by a new pressure-perfusion technique. For cortical cells in the elongating and mature regions of the hypocotyls T 1/2=0.4–15.1 s, Lp=0.2·10-5–10.0·10-5 cm s-1 bar-1 and D c=0.1·10-6–5.5·10-6 cm2 s-1. Sorption kinetics yielded a tissue diffusivity D t=0.2·10-6–0.8·10-6 cm2 s-1. The sorption kinetics include both cell-wall and cell-to-cell pathways for water transport. By comparing D c and D t, it was concluded that during swelling or shrinking of the tissue and during growth a substantial amount of water moves from cell to cell. The pressure-perfusion technique imposed hydrostatic gradients across the cortex either by manipulating the hydrostatic pressure in the xylem of hypocotyl segments or by forcing water from outside into the xylem. In segments with intact cuticle, the hydraulic conductance of the radial path (Lpr) was a function of the rate of water flow and also of flow direction. In segments without cuticle, Lpr was large (Lpr=2·10-5–20·10-5 cm s-1 bar-1) and exceeded the corticla cell Lp. The results of the pressure-perfusion experiments are not compatible with a cell-to-cell transport and can only the explained by a preferred apoplasmic water movement. A tentative explanation for the differences found in the different types of experiments is that during hydrostatic perfusion the apoplasmic path dominates because of the high hydraulic conductivity of the cell wall or a preferred water movement by film flow in the intercellular space system. For shrinking and swelling experiments and during growth, the films are small and the cell-to-cell path dominates. This could lead to larger gradients in water potential in the tissue than expected from Lpr. It is suggested that the reason for the preference of the cell-to-cell path during swelling and growth is that the solute contribution to the driving force in the apoplast is small, and tensions normally present in the wall prevent sufficiently thick water films from forming. The solute contribution is not very effective because the reflection coefficient of the cell-wall material should be very small for small solutes. The results demonstrate that in plant tissues the relative magnitude of cell-wall versus cell-to-cell transport could dependent on the physical nature of the driving forces (hydrostatic, osmotic) involved.Abbreviations and symbols D c diffusivity of the cell-to-cell pathway - D t diffusivity of the tissue - radial flow rate per cm2 of segment surface - Lp hydraulic conductivity of plasma-membrane - Lpr radial hydraulic conductance of the cortex - T 1/2 half-time of water exchange between cell and surroundings - volumetric elastic modulus  相似文献   

2.
Summary The charge-pulse technique has been used previously for the study of quasistationary processes in membranes which required only a moderate time resolution. It is shown here that a time resolution of about 400 nsec may be achieved with this technique and that it may be applied to the kinetic analysis of carrier-mediated ion transport. By this method we have studied the transport of alkali ions through optically black monoolein membranes in the presence of the ion carrier valinomycin. All three relaxation processes that are predicted by theory have been resolved. From the relaxation times and the relaxation amplitudes the rate constants for the association (k R ) and the dissociation (k D ) of the ioncarrier complex, as well as the translocation rate constants of the complex (k MS ) and the free carrier (k S ) could be obtained. For 1m Rb+ at 25° C the values arek R =3×105 m –1 sec–1,k D =2×105 sec–1,k MS =3×105 sec–1,k S =4×104 sec–1. The activation energies of the single rate constants which have been estimated from experiments at two different temperatures range between 50 and 90 kJ/mol.  相似文献   

3.
The hydraulic conductivity of the lateral walls of early metaxylem vessels (Lpx in m · s–1 · MPa–1) was measured in young, excised roots of maize using a root pressure probe. Values for this parameter were determined by comparing the root hydraulic conductivities before and after steam-ringing a short zone on each root. Killing of living tissue virtually canceled its hydraulic resistance. There were no suberin lamellae present in the endodermis of the roots used. The value of Lpx ranged between 3 · 10–7 and 35 · 10–7 m · s–1 · MPa–1 and was larger than the hydraulic conductivity of the untreated root (Lpr = 0.7 · 10–7 to 4.0 · 10–7 m · s–1 · MPa–1) by factor of 3 to 13. Assuming that all flow through the vessel walls was through the pit membranes, which occupied 14% of the total wall area, an upper limit of the hydraulic conductivity of this structure could be given(Lppm=21 · 10–7 to 250 · 10–7 m · s–1 · MPa–1). The specific hydraulic conductivity (Lpcw) of the wall material of the pit membranes (again an upper limit) ranged from 0.3 · 10–12 to 3.8 · 10–12 m2 · s–1 · MPa–1 and was lower than estimates given in the literature for plant cell walls. From the data, we conclude that the majority of the radial resistance to water movement in the root is contributed by living tissue. However, although the lateral walls of the vessels do not limit the rate of water flow in the intact system, they constitute 8–31% of the total resistance, a value which should not be ignored in a detailed analysis of water flow through roots.Abbreviatations and Symbols kwr (T 1 2/W ) rate constant (half-time) of water exchange across root (s–1 or s, respectively) - Lpcw specific hydraulic conductivity of wall material (m2 · s–1 · MPa–1) - Lppm hydraulic conductivity of pit membranes (m · s –1 · MPa–1) - Lpr hydraulic conductivity of root (m · s–1 · MPa–1) - Lpx lateralhydraulic conductivity of walls of root xylem (m · s –1 · MPa–1) This research was supported by a grant from the Bilateral Exchange Program funded jointly by the Natural Sciences and Engineering Research Council of Canada and the Deutsche Forschungsgemeinschaft to C.A.P., and by a grant from the Deutsche Forschungsgemeinschaft, Sonderforschungsbereich 137, to E.S. The expert technical help of Mr. Burkhard Stumpf and the work of Ms. Martina Murrmann and Ms. Hilde Zimmermann in digitizing chart-recorder strips is gratefully acknowledged.  相似文献   

4.
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   

5.
Summary A technique for isolating thestratum corneum from the subjacent layers of the epithelium was developed which permits studying thestratum corneum as an isolated membrane mounted between half-chambers. The method basically consists of an osmotic shock induced by immersing a piece of skin in distilled water at 50°C for 2 min. When the membrane is bathed on each surface by NaCl-Ringer's solution, its electrical resistance is 14.1±1.3 cm2 (n=10). This value is about 1/100 of the whole skin resistance in the presence of the same solution. The hydraulic filtration coefficient (L p ) measured by a hydrostatic pressure method, with identical solutions on each side of the membrane, is 8.8×10–5±1.5×10–5 cm sec–1 atm–1 (n=10) in distilled water and 9.2×10–5±1.4×10–5 cm sec–1 atm–1 (n=10) in NaCl-Ringer's solution. These values are not statistically different and are within the range of 1/80 to 1/120 of the whole skinL p . Thestratum corneum shows an amphoteric character when studied by KCl diffusion potentials at different pH's. The membrane presents an isoelectric pH of 4.6±0.3 (n=10). Above the isoelectric pH the potassium transport number is higher than the chloride transport number; below it, the reverse situation is valid. Divalent cations (Ca++ or Cu++) reduce membrane ionic discrimination when the membrane is negatively charged and are ineffective when the membrane fixed charges are protonated at low pH.  相似文献   

6.
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   

7.
The effects of anoxia on water and solute transport across excised roots of young maize plants (Zea mays L. cv. Tanker) grown hydroponically have been studied. With the aid of the root pressure probe, root pressure (Pr), root hydraulic conductivity (Lpr), and root permeability (Psr), and reflection ( sr) coefficients were measured using potassium nitrate (a typical nutrient salt) and sodium nitrate (an atypical nutrient salt) as solutes. During a period of 10–15 h, anaerobic treatment (0.0–0.2 g O2·m-3 in root medium) caused a decrease of root pressure by 0.01–0.28 MPa (by 10–80% of original root pressure) after a short transient increase. For a time period of 5 h, the decrease in the stationary root pressure was not reversible. Under anaerobic conditions, roots still behaved like osmometers and were not leaky. The root hydraulic conductivity measured in osmotic experiments (osmotic solute: NaNO3) was smaller by one to two orders of magnitude than that measured in the presence of hydrostatic gradients. Both the osmotic and hydrostatic hydraulic conductivity decreased during anaerobic treatment by 28 and 44%, respectively, at a constant reflection coefficient of the solutes ( sr=0.3–1.0). As with root pressure, changes in root permeability to water and solutes were not reversible within 5 h. Under aerobic conditions and at low external concentrations (31–59 mOsmol·kg-1), osmotic response curves were monophasic for KNO3, i.e. there was no passive uptake of solutes. Response curves became biphasic at higher concentrations (100–150 mOsmol·kg-1)- For NaNO3, response curves were biphasic at all concentrations. Presumably, this pattern was a consequence of the fact that potassium had already accumulated in the xylem. During anoxia, accumulation of potassium in the xylem was reduced, and biphasic responses were also obtained at lower potassium concentrations applied to the medium. The results are discussed in terms of a pump/leak model of the root in which anoxia affects both the active ion pumping and the permeability of the root to nutrient salts (leakage). The effects of anaerobiosis on the passive transport properties of the root (Lpr, Psr, sr) are in line with the recently proposed composite transport model of the root.Abbreviations and Symbols Ar root surface area - Lpr root hydraulic conductivity - Lprh hydrostatic hydraulic conductivity of root - Lpro osmotic hydraulic conductivity of root - Pr root pressure - Psr permeability coefficient of root - sr reflection coefficient of root The authors thank Mr. Walter Melchior for the curve-fitting program used to work out Lprh values from root pressure relaxations and Mr. Mohammad Hajirezai (Lehrstuhl für Pflanzenphysiologie, Universität Bayreuth) for making the ATP measurements. The assistance of Mrs. Libuse Badewitz in making the drawings and the technical help of Mr. Burkhard Stumpf are also gratefully acknowledged.  相似文献   

8.
Summary Ion flux relations in the unicellular marine algaAcetabularia have been investigated by uptake and washout kinetics of radioactive tracers (22Na+,42K+,36Cl and86Rb+) in normal cells and in cell segments with altered compartmentation (depleted of vacuole or of cytoplasm). Some flux experiments were supplemented by simultaneous electrophysiological recordings. The main results and conclusions about the steady-state relations are: the plasmalemma is the dominating barrier for translocation of K+ with influx and efflux of about 100 nmol·m–2·sec–1×K+ passes three- to sevenfold more easily than Rb+ does. Under normal conditions, Cl (the substrate of the electrogenic pump, which dominates the electrical properties of the plasmalemma in the resting state) shows two efflux components of about 17 and 2 mol·m–2·sec–1, and a cytoplasmic as well as vacuolar [Cl] of about 420mm ([Cl] o =529mm). At 4°C, when the pump is inhibited, both influx and efflux, as well as the cellular [Cl], are significantly reduced. Na+ ([Na+] i : about 70mm, [Na+] o : 461mm), which is of minor electrophysiological relevance compared to K+, exhibits rapid and virtually temperature-insensitive (electroneutral) exchange (two components with about 2 and 0.2 mol·m–2·sec–1 for influx and efflux). Some results with Na+ and Cl are inconsistent with conventional (noncyclic) compartmentation models: (i) equilibration of the vacuole (with the external medium) can be faster than equilibration of the cytoplasm, (ii) absurd concentration values result when calculated by conventional compartmental analysis, and (iii) large amounts of ions can be released from the cell without changes in the electrical potential of the cytoplasm. These observations can be explained by the particular compartmentation of normalAcetabularia cells (as known by electron micrographs) with about 1 part cytoplasm, 5 parts central vacuole, and 5 parts vacuolar vesicles. These vesicles communicate directly with the central vacuole, with the cytoplasm and with the external medium.  相似文献   

9.
Summary The high membrane potential ofAcetabularia (E m=–170 mV) is due to an electrogenic pump in parallel with the passive diffusion system (E d=–80 mV) which could be studied separately in the cold, when the pump is blocked. Electrical measurements under normal conditions show that the pump pathway consists of its electromotive forceE p with two elementsP 1 andP 2 in series;P 2 is shunted by a large capacitance (C p=3 mF cm–2). The nonlinear current-voltage relationship ofP 1 (light- and temperature-sensitive) could be determined separately; it reflects the properties of a carrier-mediated electrogenic pump. The value ofE p (–190 mV) indicates a stoichiometry of 21 between electrogenically transported charges and ATP. The electrical energy, normally stored inC p, compares well with the metabolic energy, stored in the ATP pool. The nonlinear current-voltage relationship ofP 2 (attributed to phosphorylating reactions) is also sensitive to light and temperature and is responsible for the region of negative conductance of the overall current-voltage relationship. The power of the pump (1 W cm–2) amounts to some percent of the total energy turnover. The high Cl fluxes (1 nmol cm–2 sec–1) and the electrical properties of the plasmalemma are not as closely related as assumed previously. For kinetic reasons, a direct and specific Cl pathway between the vacuole and outside is postulated to exist.  相似文献   

10.
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   

11.
Summary This paper reports the inhibitory effects of calmidazolium (CDZ), a calmodulin inhibitor, on electrical uncoupling by CO2. Membrane potential and coupling ratio (V 2/V1) are measured in two neighboring cells ofXenopus embryos (16 to 64 cell stage) for periods as long as 5.5 hr. Upon exposure to 100% CO2, control cells consistently uncouple even if the CO2 treatments are repeated every 15 min for 2.5 hr. CDZ (5×10–8–1×10–7 m) strongly inhibits uncoupling. The inhibition starts after 30, 50 and 60 min of treatment with 1×10–7, 7×10–8 and 5×10–8 m CDZ, respectively, is concentration-dependent and partially reversible. In the absence of CO2, CDZ also improves electrical coupling. CDZ has no significant effect on membrane potential and nonjunctional membrane resistance. These data suggest that calmodulin or a calmodulin-like protein participates in the uncoupling mechanism.  相似文献   

12.
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   

13.
We have investigated the permeability of the human red blood cell to four di-hydroxy alcohols, 1,2PD (1,2 propanediol), 1,3PD (1.3 propanediol), 1,4BD (1,4 butanediol), and 2,3BD (2,3 butanediol), and to water by using a recently developed ESR stopped-flow method which is free from artifacts found in light scattering methods. Numerical solutions of the Kedem-Katchalsky equations fit to experimental data yielded the following permeability coefficients: P1,2PD = 3.17 × 10–5 cm sec–1, p1,3pd = 1.75 × 10–5 cm sec–1, P1,4BD = 2.05 × 105 cm sec–1, P2,3BD = 7.32 × 10–5 cm sec–1. Reflection coefficients () were evaluated by comparing data fit with assumed values of = 0.6,0.8 and 1.0. In all four cases the best fit was obtained with = 1.0. Treatment of cells with PCMBS (para-chloro mercuri-benzenesulfonate) was followed by a large (> 10-fold) decrease in water permeability with virtually no change in alcohol permeability. We conclude that these alcohols do not permeate the water channels to any significant extent, and discuss some of the problems in light scattering measurements of reflection coefficients that could lead to erroneous values for .We would like to thank Professor Lenore W. Yousef (Dept. of Biology, California State Univ., Fresno) for valuable discussions and critical comments. We thank Lidia Mannuzzu for measurements of ESR spectra in the presence and absence of alcohol. We are also indebted to Kate Van Fossen for her dedicated technical support. This work was supported by NIH grant No. HL-20985.  相似文献   

14.
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   

15.
Nitrate reduction was studied in the dinoflagellatePeridinium cinctum collected from extensive algal blooms in Lake Kinneret (Israel).Among several methods tested for the preparation of cell free extracts, only the use of a ground-glass tissue culture homogenizer was found to be efficient. The assimilatory nitrate reductase ofP. cinctum was located in a particulate fraction. In this respect,P. cinctum did not behave like other eukaryotes, such as green algae, but as a prokaryote. Nitrite reductase activity was found in the soluble fraction.Nitrate reductase used NADH as a preferable electron donor; it reacted also with NADPH but only to give 16.5% of the NADH dependent rate. Methyl viologen and benzyl viologen could also serve as electron donors, with rates higher than the NADH dependent activity (3–6 times and 1.5–3 times, respectively). The Km of nitrate reductase for NADH was 2.8×10–4 M and for NO3-1.9×10–4 M. Flavins did not stimulate the activity, nor was ferricyanide able to activate it. Carboxylic anions stimulated nitrate reductase activity 3–4 fold, an effect which was not mimicked by other anions.Chlorate, azide and cyanide were competitive inhibitors ofP. cinctum, nitrate reductase withK i values of 1.79×10–3 M, 2.1×10–5 M and 8.9×10–6 M respectively.  相似文献   

16.
17.
Summary In cells of the freshwater algaHydrodictyon africanum, in solutions where [K+]0=0.1mm and pH0>7.0, the membrane in the light is hyperpolarized. The membrane potential difference {ie179-1} has values from –180 to –275 mV, more negative than any ion diffusion potential difference, and is predominantly a function of pH0, and independent of [K+]0. The hyperpolarization of the membrane appears to arise from an electrogenic efflux of H+, estimated from voltage-clamp data to be about 8 nmol m–2 sec–1 when pH0=8.5. In the light the membrane conductanceg m is about 0.084 S m–2. At light-off, {ie179-2} becomes less negative, with a halftime for change of 15 to 30 sec andg m decreases by about 0.052 S m–2. After dark periods of up to 300 sec, {ie179-3} is largely independent of pH0 for values greater than 6.0 and usually behaves as a combined K+ and Na+ diffusion potential with permeability ratioP Na/P K=0.05 to 0.2. The membrane potassium conductanceg K has either a low value of 2–6×10–2 Sm–2, or a high value of up to 18×10–2 S m–2 depending on [K+]0, the transition from low to high values occurring when {ie179-4} moves over a threshold value that is more negative than {ie179-5}, the electrochemical equilibrium potential for K+. The time for half-change of the transition is about 30 sec. The results are consistent with a model of the membrane in which the pump electromotive force and conductance are in parallel with diffusive electromotive forces and conductances. When the pump is operating its properties determine membrane properties, and when it is inoperative, or running at a diminished rate, the membrane properties are determined more by the diffusive pathways. Changes in both pump rate andg K can account for a variety of characteristic changes in membrane PD and conductance occurring in response to ligh-dark changes, changes in light intensity, pasage of externally applied electric current across the membrane and changes in ionic constituents of the external medium.  相似文献   

18.
Peptide segments derived from consensus sequences of the inhibitory site of cystatins, the natural inhibitors of cysteine proteinases, were used to develop new substrates and inhibitors of papain and rat liver cathepsins B, H, and L. Papain hydrolyzedAbz-QVVAGA-EDDnp andAbz-LVGGA-EDDnp at about the same rate, with specificity constants in the 107M–1 sec–1 range; cathepsin L also hydrolyzes both substrates with specificity constants in the 105 M–1 sec–1 range due to lowerk cat values, with theK m 's being identical to those with papain. OnlyAbz-LVGGA-EDDnp was rapidly hydrolyzed by cathepsin B, and to a lesser extent by cathepsin H. Peptide substrates that alternate these two building blocks (LVGGQVVAGAPWK and QVVAGALVGGAPWK) discriminate the activities of cathepsins B and L and papain. Cathepsin L was highly selective for cleavage at the G-G bond of the LVGG fragment in both peptides. Papain and cathepsin B cleaved either the LVGG fragment or the QVVAG fragment, depending on their position within the peptide. While papain was more specific for the segment located C-terminally, cathepsin B was specific for that in N-terminal position. Peptidyl diazomethylketone inhibitors based on these two sequences also reacted differently with papain and cathepsins. GlcA-QVVA-CHN2 was a potent inhibitor of papain and reacted with papain 60 times more rapidly (k +0= 1,100,000 M–1 sec–1) than with cathepsin L, and 220 times more rapidly than with cathepsin B. Cathepsins B and L were preferentially inhibited by Z-RLVG-CHN2. Thus cystatin-derived peptides provide a valuable framework for designing sensitive, selective substrates and inhibitors of cysteine proteinases.  相似文献   

19.
Summary Reaction kinetic analysis of the electrical properties of the electrogenic Cl pump inAcetabularia has been extended from steady-state to nonsteady-state conditions: electrical frequency responses of theAcetabularia membrane have been measured over the range from 1 Hz to 10 kHz at transmembrane potential differences across the plasmalemma (V m ) between –70 and –240 mV using voltage-clamp techniques. The results are well described by an electrical equivalent circuit with three parallel limbs: a conventional membrane capacitancec m , a steadystate conductanceg o (predominantly of the pump pathway plus a minor passive ion conductance) and a conductanceg s in series with a capacitancec p which are peculiar to the temporal behavior of the pump. The absolute values and voltage sensitivities of these four elements have been determined:c m of about 8 mF m–2 turned out to be voltage insensitive; it is considered to be normal.g o is voltage sensitive and displays a peak of about 80 S m–2 around –180 mV. Voltage sensitivity ofg s could not be documented due to large scatter ofg s (around 80 S m–2).c p behaved voltage sensitive with a notch of about 20 mF m–2 around –180 mV, a peak of about 40 mF m–2 at –120 mV and vanishing at –70 mV. When these data are compared with the predictions of nonsteady-state electrical properties of charge transport systems (U.-P. Hansen, J. Tittor, D. Gradmann, 1983,J. Membrane Biol. in press), model A (redistribution of states within the reaction cycle) consistently provides magnitude and voltage sensitivity of the elementsg o ,g s andc p of the equivalent circuit, when known kinetic parameters of the pump are used for the calculations. This analysis results in a density of pump elements in theAcetabularia plasmalemma of about 50 nmol m–2. The dominating rate constants for the redistribution of the individual states of the pump in the electric field turn out to be in the range of 500 sec–1, under normal conditions.  相似文献   

20.
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   

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