首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The reflection coefficient, sigma, for several small permeant nonelectrolytes was determined for dog and beef red blood cell membranes. Our sigma values were considerably higher than those previously reported for dog cells; e.g., out sigma urea was 87% higher than the sigma urea of Rich, Sha'afi, Barton and Solomon (J. Gen. Physiol. 50: 2391, 1967). Our sigma values for urea were only slightly greater in beef cells than previously reported by Farmer and Macey (Biochim. Biophys. Acta 290: 290, 1972). We found that a trend exists when (1 - sigma) is plotted against the log of the permeability coefficient, omega. This observation is also consistent with our previously reported sigma data for human red cell membranes (Owen & Eyring, J. Gen. Physiol. 66: 241, 1972). This trend suggests that small hydrophilic molecules interact highly with cell membrane water. The exceptions to this trend were lipophilic molecules, indicating they do not interact with water while penetrating the red cell membrane.  相似文献   

2.
We have studied the permeability of a series of hydrophilic amides and ureas through the red cell membrane by determining the three phenomenological coefficients which describe solute-membrane interaction: the hydraulic permeability (Lp), the phenomenological permeability coefficient (omega i) and the reflection coefficient (sigma i). In 55 experiments on nine solutes, we have determined that the reflection coefficient (after a small correction for solute permeation by membrane dissolution) is significantly less than 1.0 (P less than 0.003, t-test), which provides very strong evidence that solute and water fluxes are coupled as they cross the red cell membrane. It is proposed that the aqueous channel is a tripartite assembly, comprising H-bond exchange regions at both faces of the membrane, joined by a narrower sieve-specific region which crosses the lipid. The solutes bind to the H-bond exchange regions to exchange their solvation shell with the H-bonds of the channel; the existence of these regions is confirmed by the finding that the permeation of all the amides and ureas requires binding to well-characterized sites with Km values of 0.1-0.5 M. The sieve-specific regions provide the steric restraints which govern the passage of the solutes according to their size; their existence is shown by the findings that: (1) the reflection coefficient (actually the function [1-corrected sigma i]) is linearly dependent upon the solute molecular diameter; and (2) the permeability coefficient is linearly dependent upon solute molar volume. These several observations, taken together, provide strong arguments which lead to the conclusion that the amides and urea cross the red cell membrane in an aqueous pore.  相似文献   

3.
This study presents a generic numerical model to simulate the coupled solute and solvent transport in human ovarian tissue sections during addition and removal of chemical additives or cryoprotective agents (CPA). The model accounts for the axial and radial diffusion of the solute (CPA) as well as axial convection of the CPA, and a variable vascular surface area (A) during the transport process. In addition, the model also accounts for the radial movement of the solvent (water) into and out of the vascular spaces. Osmotic responses of various cells within an human ovarian tissue section are predicted by the numerical model with three model parameters: permeability of the tissue cell membrane to water (L(p)), permeability of the tissue cell membrane to the solute or CPA (omega) and the diffusion coefficient of the solute or CPA in the vascular space (D). By fitting the model results with published experimental data on solute/water concentrations within an human ovarian tissue section, I was able to determine the permeability parameters of ovarian tissue cells in the presence of 1.5M solutions of each of the following: dimethyl sulphoxide (DMSO), propylene glycol (PROH), ethylene glycol (EG), and glycerol (GLY), at two temperatures (4 degrees C and 27 degrees C). Modeling Approach 1: Assuming a constant value of solute diffusivity (D = 1.0 x 10(-9) m(2)/sec), the best fit values of L(p) ranged from 0.35 x 10(-14) to 1.43 x 10(-14) m(3)/N-sec while omega ranged from 2.57 x 10(-14) to 70.5 x 10(-14) mol/N-sec. Based on these values of L(p) and omega, the solute reflection coefficient, sigma defined as sigma = 1-omega v(CPA)/L(P) ranged from 0.9961 to 0.9996. Modeling Approach 2: The relative values of omega and sigma from our initial modeling suggest that the embedded ovarian tissue cells are relatively impermeable to all the CPAs investigated (or omega approximately 0 and sigma approximately 1.0). Consequently the model was modified and used to predict the values of L(p) and D assuming omega = 0 and sigma = 1.0. The best fit values of L(p) ranged from 0.44 x 10(-14) to 1.2 x 10(-14) m(3)/N-sec while D ranged from 0.85 x 10(-9) to 2.08 x 10(-9) m(2)/sec. Modeling Approach 3: Finally, the best fit values of D from modeling approach 2 were incorporated into model 1 to re-predict the values of L(p) and omega. It is hoped that the ovarian tissue cell parameters reported here will help to optimize chemical loading and unloading procedures for whole ovarian tissue sections and consequently, tissue cryopreservation procedures.  相似文献   

4.
Quantum mechanical calculations of transmission coefficients for some permeant molecules across the human red cell and resting axolemma squid axon membranes are carried out. The calculations depend on (i) the molecular weight of the molecule and (ii) the depth and width of the potential well of the membrane. In most cases good agreement between calculated and experimental values is found.  相似文献   

5.
6.
Stoichiometry of subunits in the H+-ATPase complex of Escherichia coli   总被引:35,自引:0,他引:35  
The H+-ATPase (F1F0) of Escherichia coli was purified from cells labeled with either [35S]sulfate or [U-14C-D] glucose, and the molar ratio of subunits in the complex determined. The molar ratio was calculated from the radioactivity incorporated into each subunit, using either the subunit sulfur content or subunit molecular weight. These labeling experiments confirm an alpha 3 beta 3 gamma 1 delta 1 epsilon 1 ratio of subunits in F1, and indicate a chi 1 psi 2 omega 10 ratio of subunits in F0. The chi, psi, and omega designations used here refer to the subunits of F0 in order of decreasing molecular weight. Staining with Coomassie brilliant blue gave a reliable indication of the molar ratio of subunits in F1, but very erroneous values for each of the subunits of F0. We attempted to estimate the ratio of subunits in the native membrane, since the stoichiometry determined for the purified complex could be an anomaly of purification. These estimates were made after labeling cells with [35S]sulfate during amplification of the ATPase genes carried on a lambda transducing phage. The subunit ratios in the native membrane were reasonably close to those obtained with purified F1F0. We conclude that the stoichiometry determined reflects the composition of F1F0 in the native membrane. The most surprising conclusion from this study is that there are 10 +/- 1 omega ("proteolipid") subunits in each F1F0 complex. This is considerably more than had been assumed previously.  相似文献   

7.
Solute interactions with membrane proteins can be analyzed by biomembrane affinity chromatography (BAC), previously applied to the human red cell glucose transporter. As a novel example, frontal BAC analysis of interactions between the nucleoside transport inhibitor nitrobenzylthioinosine (NBTI) and immobilized reconstituted nucleoside and glucose transporters from human red cells revealed two binding sites, presumably corresponding to the two transporters. The affinities and amounts of sites were determined by use of a double rectangular hyperbolic equation. The Kd value for NBTI binding to the nucleoside transporter in egg phospholipid proteoliposomes was 0.38 +/- 0.08 nM (22 degrees C, I = 0.16, pH 7.4), lower than previously reported for reconstituted systems. The molar ratio between the amounts of nucleoside transporter sites for NBTI and glucose transporter sites for cytochalasin B was 4.5 +/- 0.6%.  相似文献   

8.
Monocytes were isolated from fresh whole human blood and resuspended in Hanks balanced salt solution; a portion of the cells was mixed with an equal volume of 2M dimethyl sulfoxide (DMSO) to form a 1 M solution. Microliter volumes of cell suspension were placed directly onto a computer-controlled cryostage and cooled to a predetermined subzero temperature. Ice was nucleated in the extracellular medium and a continuous video record was made of the subsequent osmotically induced volume changes of individual cells owing to exposure to the concentrated extracellular solutes. Selected micrographs emphasizing the initial transient data were digitized for computer analysis with an interactive boundary tracing algorithm to determine metric parameters of specific cells, and apparent volume changes were measured as a function of elapsed time after nucleation. The Kedem-Katchalsky-coupled transport equations were fit to the data using a network thermodynamic model implemented on a microcomputer to determine values for the permeability properties Lp, omega, and sigma. Experiments were performed over the temperature range from -7 degrees to -10 degrees C. Cells pre-equilibrated with DMSO had a lower Lp and a higher activation energy, delta E, than without additive, although the statistical significance of the difference could not be substantiated. It was found that the movement of DMSO across the plasma membrane in response to extracellular freezing was apparently so much smaller than the water flux that values for omega and sigma could not be determined from the data base.  相似文献   

9.
BACKGROUND: Experimental data and a complementary biophysical model are presented to describe the dynamic response of a unicellular microalga to osmotic processes encountered during cryopreservation. METHOD OF APPROACH: Chlorococcum texanum (C. texanum) were mounted on a cryoperfusion microscope stage and exposed sequentially to various solutions of sucrose and methanol. Transient volumetric excursions were determined by capturing images of cells in real time and utilizing image analysis software to calculate cell volumes. A biophysical model was applied to the data via inverse analysis in order to determine the plasma membrane permeability to water and to methanol. The data were also used to determine the elastic modulus of the cell wall and its effect on cell volume. A three-parameter (hydraulic conductivity (Lp), solute permeability; (omega), and reflection coefficient, (sigma)) membrane transport model was fit to data obtained during methanol perfusion to obtain constitutive property values. These results were compared with the property values obtained for a two coefficient (Lp and omega) model. RESULTS: The three-parameter model gave a value for sigma not consistent with practical physical interpretation. Thus, the two-coefficient model is the preferred approach for describing simultaneous water and methanol transport. The rate of both water and methanol transport were strongly dependent on temperature over the measured temperature range (25 degrees C to -5 degrees C) and cells were appreciably more permeable to methanol than to water at all measured temperatures. CONCLUSION: These results may explain in part why methanol is an effective cryoprotective agent for microalgae.  相似文献   

10.
Solute reflection coefficients sigma of cell membrane vesicles or liposomes are commonly determined by comparison of the water flow induced by a gradient of the studied solute and that of a reference molecule using light scattering techniques. However, variations in scattered light which are mainly related to change in vesicle volume are also influenced by the refractive index of the surrounding medium. Therefore comparing kinetics of vesicle shrinkage induced by hyperosmotic solutions which have different refractive indexes might lead to an under or over estimation of sigma. We determined sigma NaCl in rat kidney brush-border membrane vesicles by two different approaches using mannitol, a poorly permeant molecule, as reference. (1) The refractive index of the hyperosmotic NaCl solution was adjusted to that of mannitol by addition of polyvinyl pyrrolidone (Mr 40,000), without a significant increase in osmolality. Thereby the change in scattered light intensity induced by osmotic vesicle shrinkage due to gradients of NaCl and mannitol were comparable and led to a sigma NaCl value close to one instead of the previously published value of 0.53. (2) The reflection coefficient was calculated from the lifetime of vesicle shrinkage which is not refractive index-dependent. Again sigma NaCl was not different from one. These results suggest that the water proteic pathways found in the luminal membrane of proximal tubules are not shared by salts.  相似文献   

11.
Quantum mechanical calculations of transmission coefficients for some permeant molecules across the human red cell and resting axolemma squid axon membranes are carried out. The calculations depend on (i) the molecular weight of the molecule and (ii) the depth and width of the potential well of the membrane. In most cases good agreement between calculated and experimental values is found.  相似文献   

12.
The rate of exchange of urea across the membranes of human erythrocytes (red blood cells) was quantified on the 1-s to 2-min timescale. 13C-urea was hyperpolarized and subjected to rapid dissolution and the previously reported (partial) resolution of 13C NMR resonances from the molecules inside and outside red blood cells in suspensions was observed. This enabled a stopped-flow type of experiment to measure the (initially) zero-trans transport of urea with sequential single-pulse 13C NMR spectra, every second for up to ∼2 min. Data were analyzed using Bayesian reasoning and a Markov chain Monte Carlo method with a set of simultaneous nonlinear differential equations that described nuclear magnetic relaxation combined with transmembrane exchange. Our results contribute to quantitative understanding of urea-exchange kinetics in the whole body; and the methodological approach is likely to be applicable to other cellular systems and tissues in vivo.  相似文献   

13.
The rate of exchange of urea across the membranes of human erythrocytes (red blood cells) was quantified on the 1-s to 2-min timescale. 13C-urea was hyperpolarized and subjected to rapid dissolution and the previously reported (partial) resolution of 13C NMR resonances from the molecules inside and outside red blood cells in suspensions was observed. This enabled a stopped-flow type of experiment to measure the (initially) zero-trans transport of urea with sequential single-pulse 13C NMR spectra, every second for up to ∼2 min. Data were analyzed using Bayesian reasoning and a Markov chain Monte Carlo method with a set of simultaneous nonlinear differential equations that described nuclear magnetic relaxation combined with transmembrane exchange. Our results contribute to quantitative understanding of urea-exchange kinetics in the whole body; and the methodological approach is likely to be applicable to other cellular systems and tissues in vivo.  相似文献   

14.
3,5,3'-Triiodo-L-thyronine (L-T3)-binding sites from rat and human red cells were characterized as to their distribution between the two surfaces of the membrane. Analysis of L-T3 binding to sealed right-side-out and inside-out vesicles from erythrocyte membrane revealed that high affinity L-T3-binding sites are located on the external side in rat erythrocytes and on the internal side in human red cells. These results were further confirmed by preincubation of intact red cells with p-chloromercuribenzoate, a slowly permeant reagent that interacts reversibly with SH groups of proteins. Following this treatment only the SH groups of L-T3 sites from rat erythrocytes were found to be blocked. Scatchard analysis of the binding data for rat right-side-out and human inside-out vesicles showed high affinity sites with Kd values of 0.2 x 10(-10) and 2 x 10(-10) M, respectively. The results suggest that the orientation of L-T3-binding sites in the erythrocyte membrane is species-dependent.  相似文献   

15.
The first crystal structure of human telomeric DNA in complex with the natural alkaloid berberine, produced by different plant families and used in folk medicine for millennia, was solved by X-ray diffraction method. The G-quadruplex unit features all-parallel strands. The overall folding assumed by DNA is the same found in previously reported crystal structures. Similarly to previously reported structures the ligand molecules were found to be stacked onto the external 5′ and 3′-end G-tetrads. However, the present crystal structure highlighted for the first time, the presence of two berberine molecules in the two binding sites, directly interacting with each tetrad. As a consequence, our structural data point out a 2:1 ligand to G-tetrad molar ratio, which has never been reported before in a telomeric intramolecular quadruplex structure.  相似文献   

16.
Liu J  Christian JA  Critser JK 《Cryobiology》2002,44(3):2409-268
The objective of this study was to determine the cryobiological characteristics of canine red blood cells (RBC). These included the hydraulic conductivity (L(p)), the permeability coefficients (P(s)) of common cryoprotectant agents (CPAs), the associated reflection coefficient (sigma), the activation energies (E(a)) of L(p) and P(s) and the osmotic tolerance limits. By using a stopped-flow apparatus, the changes of fluorescence intensity emitted by intracellularly entrapped 5-carboxyfluorescein diacetate (CFDA) were recorded when cells were experiencing osmotic volume changes. After the determination of the relationship between fluorescence intensity and cell volume, cell volume changes were calculated. These volume changes were used in three-parameter fitting calculations to determine the values of L(p), P(s), and sigma for common CPAs. These volume measurements and data analyses were repeated at three different temperatures (22, 14, 7 degrees C). Using the Arrhenius equation, the activation energies of L(p) and P(s) in the presence of CPAs were determined. The osmotic tolerance limits for canine RBC were determined by measuring the percentage of free hemoglobin in NaCl solutions with various osmolalities compared to that released by RBC incubated in double distilled water. The upper and lower osmotic tolerance limits were found to be 150mOsm (1.67V(iso)) and 1200mOsm (0.45V(iso)), respectively. These parameters were then used to calculate the amount of non-permeating solute needed to keep cell volume excursions within the osmotic tolerance limits during CPA addition and removal.  相似文献   

17.
A stopped-flow rapid-reaction apparatus was used to follow the time course of extracellular pH in a human red cell suspension following a sudden increase in PCO2. The extracellular pH change was slow (t1/2 similar to 3.5 s) considering the presence of carbonic anhydrase in the cells. When carbonic anhydrase was added to the extracellular fluid, the half-time was reduced to less than 20 ms. The explanation for these phenomena is that the equilibration of H+ across the red cell membrane is rate-limited by the uncatalyzed reaction CO2 plus H2O formed from H2CO3 outside the cells. A theoretical model was developed which successfully reproduced the experimental results. When the model was used to simulate CO2 exchange in vivo, it was determined that blood PCO2 and pH require long times (greater than 50 s) to approach equilibrium between cells and plasma after leaving an exchange capillary. We conclude that cell-plasma equilibrium may never be reached in vivo, and that in vitro measurements of these quantities may not represent their true values at the site of sampling.  相似文献   

18.
Solubility-diffusion theory, which treats the lipid bilayer membrane as a bulk lipid solvent into which permeants must partition and diffuse across, fails to account for the effects of lipid bilayer chain order on the permeability coefficient of any given permeant. This study addresses the scaling factor that must be applied to predictions from solubility-diffusion theory to correct for chain ordering. The effects of bilayer chemical composition, temperature, and phase structure on the permeability coefficient (Pm) of acetic acid were investigated in large unilamellar vesicles by a combined method of NMR line broadening and dynamic light scattering. Permeability values were obtained in distearoylphosphatidylcholine, dipalmitoylphosphatidylcholine, dimyristoylphosphatidylcholine, and dilauroylphosphatidylcholine bilayers, and their mixtures with cholesterol, at various temperatures both above and below the gel-->liquid-crystalline phase transition temperatures (Tm). A new scaling factor, the permeability decrement f, is introduced to account for the decrease in permeability coefficient from that predicted by solubility-diffusion theory owing to chain ordering in lipid bilayers. Values of f were obtained by division of the observed Pm by the permeability coefficient predicted from a bulk solubility-diffusion model. In liquid-crystalline phases, a strong correlation (r = 0.94) between f and the normalized surface density sigma was obtained: in f = 5.3 - 10.6 sigma. Activation energies (Ea) for the permeability of acetic acid decreased with decreasing phospholipid chain length and correlated with the sensitivity of chain ordering to temperature, [symbol: see text] sigma/[symbol: see text](1/T), as chain length was varied. Pm values decreased abruptly at temperatures below the main phase transition temperatures in pure dipalmitoylphosphatidylcholine and dimyristoylphosphatidylcholine bilayers (30-60-fold) and below the pretransition in dipalmitoylphosphatidylcholine bilayers (8-fold), and the linear relationship between in f and sigma established for liquid-crystalline bilayers was no longer followed. However, in both gel and liquid-crystalline phases in f was found to exhibit an inverse correlation with free surface area (in f = -0.31 - 29.1/af, where af is the average free area (in square angstroms) per lipid molecule). Thus, the lipid bilayer permeability of acetic acid can be predicted from the relevant chain-packing properties in the bilayer (free surface area), regardless of whether chain ordering is varied by changes in temperature, lipid chain length, cholesterol concentration, or bilayer phase structure, provided that temperature effects on permeant dehydration and diffusion and the chain-length effects on bilayer barrier thickness are properly taken into account.  相似文献   

19.
It had been previously reported that aromatic amines were not substrates for the bacterial quinoprotein methylamine dehydrogenase. In this study, benzylamine-dependent activity was also not observed in the steady-state assay of this enzyme with the artificial electron acceptor phenazine ethosulfate (PES). Benzylamines did, however, stoichiometrically reduce the protein-bound tryptophan tryptophylquinone (TTQ) prosthetic group and acted as reversible competitive inhibitors of methylamine oxidation when the enzyme was assayed with PES. When methylamine dehydrogenase activity was monitored using a steady-state assay which employed its physiological electron acceptor amicyanin instead of PES, very low but detectable benzylamine-dependent activity was observed. The reactions of a series of para-substituted benzylamines with methylamine dehydrogenase were examined. A Hammett plot of the log of Ki values for the competitive inhibition by these amines against sigma p exhibited a negative slope. Rapid kinetic measurements allowed the determination of values of k3 and Ks for the reduction of TTQ by each of these amines. A Hammett plot of log k3 versus sigma p exhibited a positive slope, which suggests that the oxidation of these amines by methylamine dehydrogenase proceeds through a carbanionic reaction intermediate. A negative slope was observed for the correlation between log Ks and sigma p. Plots of log k3 and log Ks against substituent constants which reflected either resonance or field/inductive parameters for each para substituent indicated that the magnitude of k3 was primarily influenced by field/inductive effects while Ks was primarily influenced by resonance effects. No correlation was observed between either k3 or Ks and the relative hydrophobicity of the para-substituted benzylamines or steric parameters.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
A previously clonedpdxHgene fromEscherichia colicoding for pyridoxine 5′-phosphate oxidase was transferred to a pET22b vector and expressed inE. coliHMS174(DE3) cells. The soluble overexpressed enzyme was rapidly purified in high yield using two chromatography columns with an overall purification of about 2.8-fold. The purified enzyme contained tightly bound FMN. The enzyme exhibited the same spectral properties and similar kinetic constants to those previously reported by G. Zhao and M. E.Winkler (J. Bacteriol.177, 883, 1995), but differed from the properties reported by other investigators. A rapid procedure was developed for preparing apoPNP Ox in high yield. Both the holo- and apoenzymes were homodimers. The molar absorbtivity coefficient for the protein was determined for the fully active apoPNP Ox from is amino acid composition. Using this value and the spectral properties of the bound FMN it was shown by three different methods that the dimeric enzyme contains two molecules of bound FMN per dimer and not one FMN as previously reported.  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