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.
We studied intracellular binding and possible compartmentalization of the fluorescent Ca2+ indicators, indo-1 and fura-2, in single mammalian cardiac ventricular cells that had been loaded with indo-1 and fura-2 by exposure to the acetoxymethylester form of the indicators (indo-1/AM and fura-2/AM). Techniques similar to those used in experiments on fluorescence recovery after photobleaching (FRAP) were used. It was assumed that reversible binding in myoplasm would be evident as slowed recovery of fluorescence after photobleaching, and that irreversible binding of the indicators to immobile myoplasmic sites (or "compartmentalization" in organelles) would be evident as incomplete recovery. Through the use of a mask, one half of a cell was exposed to high-intensity ultraviolet (UV) light to bleach the indo-1 or fura-2 in only that part of the cell. Upon removal of the mask and termination of the high-intensity UV illumination, fluorescence recovered in the bleached half of the cell, indicating diffusion of indo-1 and fura-2. Mathematical modeling of the diffusional redistribution of the indicators indicated that in these cells the apparent diffusion coefficient for indo-1 is 1.57 x 10(-7) cm2 s-1 (SD 0.48 x 10(-7) cm2 s-1; n = 5 cells, 21 degrees C), and for fura-2 is 3.19 x 10(-7) cm2 s-1 (SD 1.85 x 10(-7) cm2 s-1; n = 6 cells, 21 degrees C). These values are approximately 6 and 3, respectively, times smaller than those expected for free diffusion in the myoplasm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
We measured the lateral diffusion of the fluorescent lipid analogue dioctadecylindocarbocyanine iodide (DiI) and of membrane glycoproteins labeled with tetramethylrhodamine (TRITC) succinyl concanavalin A (SConA) via fluorescence photobleaching recovery (FPR) at selected times during a temperature downshift experiment on transformation-defective temperature-sensitive (td-ts) Rous sarcoma virus (RSV) NY68-transformed chicken embryo fibroblasts (CEF) and on identically treated CEF and RSV-transformed CEF. There were no significant differences in the lateral diffusion in DiI at any of the times measured. The lateral diffusion of TRITC-SConA on the RSV-transformed CEF, (1.32 +/- 0.12).10(-10) cm2 s-1, was approximately two times faster than that observed in normal CEF, (0.61 +/- 0.06).10(-10) cm2 s-1. In the cells undergoing RSV NY68-mediated transformation, TRITC-SConA diffusion increased over a 24-h period from a value comparable to that observed in normal CEF, (0.72 +/- 0.13).10(-10) cm2 s-1 to a value comparable to the RSV-CEF transformed cells, (1.74 +/- 0.20).10(-10) cm2 s-1. All diffusion measurements reported were made at the permissive temperature for RSV-NY68 (35 degrees C) unless stated otherwise. The changes in the lateral diffusion of TRITC-SConA occurred between the fifth and twelfth hour of the downshift course and could be associated with cytoskeletal disruption and/or fibronectin degradation, both known to occur at this time in RSV-transformed cells. To assess the contribution of extracellular matrix (ECM) degradation, SConA mobility was measured in normal and RSV-transformed cells treated with trypsin. This treatment increased SConA mobility approximately 4-fold in the normal cells relative to untreated controls and only 2-fold in the RSV-CEF transformed cells. No significant difference in SConA mobility between trypsinized spherical normal and transformed cells was apparent.  相似文献   

3.
The characteristics of water diffusional permeability (P) of human red blood cells were studied on isolated erythrocytes and ghosts by a doping nuclear magnetic resonance technique. In contrast to all previous investigations, systematic measurements were performed on blood samples obtained from a large group of donors. The mean values of P ranged from 2.2 X 10(-3) cm.s-1 at 5 degrees C to 8.1 X 10(-3) cm.s-1 at 42 degrees C. The reasons for some of the discrepancies in the permeability coefficients reported by various authors were found. In order to estimate the basal permeability, the maximal inhibition of water diffusion was induced by exposure of red blood cells to p-chloromercuribenzenesulfonate (PCMBS) under various conditions (concentration, duration, temperature). The lowest values of P were around 1.3 X 10(-3) cm.s-1 at 20 degrees C, 1.6 X 10(-3) cm.s-1 at 25 degrees C, 1.9 X 10(-3) cm.s-1 at 30 degrees C and 3.2 X 10(-3) cm.s-1 at 37 degrees C. The results reported here represent the largest series of determinations of water diffusional permeability of human red blood cells (without or with exposure to mercurials) available in the literature, and consequently the best estimates of the characteristics of this transport process. The values of P can be taken as references for the studies of water permeability in various cells or in pathological conditions.  相似文献   

4.
Permeability of human granulocytes to water: rectification of osmotic flow   总被引:1,自引:0,他引:1  
The permeability of human granulocytes to water was studied for samples of that population with an electronic particle counter. Changes in volume were monitored as these samples were introduced suddenly into hypo- and hypertonic osmotica. Temperature and concentration sensitivity analyses of the permeability coefficients were carried out. An apparent rectification of water flow seems to occur with the water permeability being three times higher for endosmotic flow than for exosmotic flow. The estimated water permeability at the reference temperature 20 degrees C was for exosmotic flow 2.04 +/- 0.02 x 10(-12) kgmol N-1 s-1 with an apparent transfer energy of 4.76 +/- 0.09 x 10(7) J kgmol-1, and for endosmotic flow 6.05 +/- 0.07 x 10(-12) kgmol N-1 s-1 with an apparent transfer energy of 4.9 +/- 0.1 x 10(7) J kgmol-1.  相似文献   

5.
Electrical resistance of muscle capillary endothelium.   总被引:2,自引:0,他引:2       下载免费PDF全文
A recently developed technique for in vivo determination of the electrical resistance of vascular endothelium in microvessels was applied to the vessels in a thin frog muscle, m. cutaneus pectoris. The technique consists of injection of current via a glass micropipette into a capillary and measurement of the resulting intra- and extravascular potential profiles with another micropipette placed at various distances from the current source. The theory of Peskoff and Eisenberg (1974) was used to handle the problems arising from distributed extravascular resistances and was experimentally shown to describe the external field satisfactorily. With this extension of one-dimensional cable theory the specific electrical resistance of arterial microvessels was 33 omega cm2 and of venous capillaries 23 omega cm2. The "length constants" were 135 and 112 micrometers, respectively. If results from arterial and venous vessels are taken together, the ionic permeabilities at 20 degrees C were PNa = 3.9 X 10(-5) cm X s-1, PK = 5.7 X 10(-5) cm X s-1, PCl = 5.9 X 10(-5) cm X s-1 and PHCO3 = 3.4 X 10(-5) cm X s-1. These figures agree with figures for capillary permeability obtained in tracer experiments on whole muscle. The study bridges a gap between single capillary and whole organ techniques with the conclusion that the two different approaches lead to similar results in muscle capillaries.  相似文献   

6.
Homotypic adhesion o2 neutrophils stimulated with chemoattractant is analogous to capture on vascular endothelium in that both processes depend on L-selectin and beta 2-integrin adhesion receptors. Under hydrodynamic shear, cell adhesion requires that receptors bind sufficient ligand over the duration of intercellular contact to withstand hydrodynamic stresses. Using cone-plate viscometry to apply a uniform linear shear field to suspensions of neutrophils, we conducted a detailed examination of the effect of shear rate and shear stress on the kinetics of cell aggregation. A collisional analysis based on Smoluchowski's flocculation theory was employed to fit the kinetics of aggregation with an adhesion efficiency. Adhesion efficiency increased with shear rate from approximately 20% at 100 s-1 to approximately 80% at 400 s-1. The increase in adhesion efficiency. Adhesion efficiency increased with shear rate from approximately 20% at 100 s-1 to approximately 80% at 400 s-1. The increase in adhesion efficiency with shear was dependent on L-selectin, and peak efficiency was maintained over a relatively narrow range of shear rates (400-800 s-1) and shear stresses (4-7 dyn/cm2). When L-selectin was blocked with antibody, beta 2-integrin (CD11a, b) supported adhesion at low shear rates (< 400 s-1). The binding kinetics of selectin and integrin appear to be optimized to function within discrete ranges of shear rate and stress, providing an intrinsic mechanism for the transition from neutrophil tethering to stable adhesion.  相似文献   

7.
Reductive titration curves of flavodoxin from Desulfovibrio vulgaris displayed two one-electron steps. The redox potential E-2 for the couple oxidized flavodoxin/flavodoxin semiquinone was determined by direct titration with dithionite. E-2 was -149 plus or minus 3 mV (pH 7.78, 25 degrees C). The redox potential E-1 for the couple flavodoxin semiquinone/fully reduced flavodoxin was deduced from the equilibrium concentration of these species in the presence of hydrogenase and H-2. E-1 was -438 plus or minus 8 mV (pH 7.78, 25 degrees C). Light-absorption and fluorescence spectra of flavodoxin in its three redox states have been recorded. Both the rate and extent of reduction of flavodoxin semiguinone with dithionite were found to depend on pH. An equilibrium between the semiquinone and hydroquinone forms occurred at pH values close to the neutrality, even in the presence of a large excess of dithionite, suggesting an ionization in fully reduced flavodoxin with a pK-a = 6.6. The association constants K for the three FMN redox forms with the apoprotein were deduced from the value of K (K = 8 times 10-7 M-1) measured with oxidized EMN at pH 7.0. Oxidized flavodoxin was found to comproportionate with the fully reduced protein (k-comp = 4.3 times 10-3 M-1 times s-1, pH 9.0, 22 degrees C) and with reduced free FMN (K-comp = 44 M-1 times s-1, pH 8.1, 20 degrees C). Fast oxidation of reduced flavodoxin occurred in the presence of O-2. Slower oxidation of semiquinone was dependent on pH in a drastic way.  相似文献   

8.
Activated neutrophils aggregate in a shear field via bonding of L-selectin to P-selectin glycoprotein ligand-1 (PSGL-1) followed by a more stable bonding of LFA-1 (CD11a/CD18) to intercellular adhesion molecule 3 (ICAM-3) and Mac-1 (CD11b/CD18) to an unknown counter receptor. Assuming that the Mac-1 counter receptor is ICAM-3-like in strength and number, rate processes were deconvoluted from neutrophil homoaggregation data for shear rates (G) of 100-3000 s-1 with a two-body hydrodynamic collision model (. Biophys. J. 73:2819-2835). For integrin-mediated aggregation (characteristic bond strength of 5 microdynes) in the absence of L-selectin contributions, an average forward rate of kf = 1.57 x 10(-12) cm2/s predicted the measured efficiencies for G = 100-800 s-1. For a selectin bond formation rate constant equal to the integrin bond formation rate constant, the colloidal stability of unactivated neutrophils was satisfied for a reverse rate of the L-selectin-PGSL bond corresponding to an average bond half-life of 10 ms at a characteristic bond strength of 1 microdyne. Colliding neutrophils initially bridged by at least one L-selectin-PSGL-1 bond were calculated to rotate from 8 to 50 times at G = 400 to 3000 s-1, respectively, before obtaining mechanical stability in sheared fluid of either 0.75 or 1.75 cP viscosity. Thus for G > 400 s-1, the interaction time needed for the rotating aggregates to become stable was relatively constant at 52.5 +/- 8.5 ms, largely independent of shear rate or shear stress. Aggregation data and the colloidal stability criterion can provide a consistent set of forward and reverse rate constants and characteristic bond strengths for a known time-dependent stoichiometry of receptors on cells interacting in a shear flow field.  相似文献   

9.
Mechanism of cryoprotection by extracellular polymeric solutes.   总被引:1,自引:1,他引:0       下载免费PDF全文
To elucidate the means by which polymer solutions protect cells from freezing injury, we cooled human monocytes to -80 degrees C or below in the presence of various polymers. Differential scanning calorimetric studies showed that those polymers which protect cells best have a limiting glass transition temperature (T'g) of approximately -20 degrees C; those with a T'g significantly higher or lower did not protect. Freeze-etch electron micrographs indicated that intracellular ice crystals had formed during this freezing procedure, but remained smaller than approximately 300 nm in the same proportion of cells as survived rapid thawing. We propose that cryoprotection of slowly frozen monocytes by polymers is a consequence of a T'g of -20 degrees C in the extracellular solution. In our hypothesis, the initial concentration and viscosity of protective polymer solutions reduce the extent and rate of cell water loss to extracellular ice and limit the injurious osmotic stress, which cells face during freezing at moderate rates to -20 degrees C. Below -20 degrees C, glass formation prevents further osmotic stress by isolating cells from extracellular ice crystals, virtually eliminating cell water loss at lower temperatures. On the other hand, the protective polymer solutions will allow some diffusion of water away from cells at temperatures above T'g. If conditions are correct, cells will concentrate the cytoplasm sufficiently during the initial cooling to T'g to avoid lethal intracellular freezing between T'g and the intracellular Tg, which has been depressed to low temperatures by that concentration. Thus, when polymers are used as cryoprotective agents, cell survival is contingent upon maintenance of osmotic stress within narrow limits.  相似文献   

10.
Microtubule assembly kinetics. Changes with solution conditions.   总被引:1,自引:0,他引:1       下载免费PDF全文
The assembly kinetics of microtubule protein are altered by ionic strength, temperature and Mg2+, but not by pH. High ionic strength (I0.2), low temperature (T less than 30 degrees C) and elevated Mg2+ (greater than or equal to 1.2 mM) induce a transition from biphasic to monophasic kinetics. Comparison of the activation energy obtained for the fast biphasic step at low ionic strength (I0.069) shows excellent agreement with the values obtained at high ionic strength, low temperature and elevated Mg2+. From this observation it can be implied that the tubulin-containing reactant of the fast biphasic event is also the species that elongates microtubules during monophasic assembly. Second-order rate constants for biphasic assembly are 3.82(+/- 0.72) x 10(7) M-1.s-1 and 5.19(+/- 1.25) x 10(6) M-1.s-1, and for monophasic assembly the rate constant is 2.12(+/- 0.56) x 10(7) M-1.s-1. The microtubule number concentration is constant during elongation of microtubules for biphasic and monophasic assembly.  相似文献   

11.
Kinetic and equilibrium studies on the interaction of steroids with human corticosteroid-binding globulin (CBG, transcortin) were performed with pH, temperature, and steroid structure as variables. Dissociation rate constants were determined fluorometrically; the values for cortisol, corticosterone, deoxycorticosterone, and progesterone are 0.031, 0.047, 0.10, and 0.16 s-1, respectively, at 20 degrees C, pH 7.4. The pH dependence of the dissociation rate constant for the corticosterone complex below pH 10.5 at 20 degrees C is given by koff = 0.043 (1 + [H+]/10(-6.50)) s-1; above pH 11, koff = 0.030 (1 + 10(-12.15/[H+] s-1. A temperature-dependence study of koff for the cortisol and progesterone complexes gave values of 0.0028 s-1 and 0.012 s-1 at 4 degrees C, respectively, and 0.88 s-1 and 4.5 s-1 at 37 degrees C, with progesterone dissociating about four to five times faster over the entire temperature range. The affinity constants, determined by equilibrium dialysis, for the binding of cortisol, corticosterone, and progesterone at 4 degrees C were 7.9, 7.2, and 7.0 X 10(8) M-1; values of 0.40 and 0.26 X 10(8) M-1 were determined at 37 degrees C for cortisol and progesterone. The close similarity of the affinity constants of the three steroids combined with differing dissociation rates implies that the association rate changes with steroid structure, in contrast to our earlier findings with progesterone-binding globulin.  相似文献   

12.
Urea permeability of human red cells   总被引:5,自引:1,他引:4       下载免费PDF全文
The rate of unidirectional [14C]urea efflux from human red cells was determined in the self-exchange and net efflux modes with the continuous flow tube method. Self-exchange flux was saturable and followed simple Michaelis-Menten kinetics. At 38 degrees C the maximal self-exchange flux was 1.3 X 10(-7) mol cm-2 s-1, and the urea concentration for half-maximal flux, K1/2, was 396 mM. At 25 degrees C the maximal self-exchange flux decreased to 8.2 X 10(-8) mol cm-2 s-1, and K1/2 to 334 mM. The concentration-dependent urea permeability coefficient was 3 X 10(-4) cm s-1 at 1 mM and 8 X 10(-5) cm s-1 at 800 mM (25 degrees C). The latter value is consonant with previous volumetric determinations of urea permeability. Urea transport was inhibited competitively by thiourea; the half-inhibition constant, Ki, was 17 mM at 38 degrees C and 13 mM at 25 degrees C. Treatment with 1 mM p-chloromercuribenzosulfonate inhibited urea permeability by 92%. Phloretin reduced urea permeability further (greater than 97%) to a "ground" permeability of approximately 10(-6) cm s-1 (25 degrees C). This residual permeability is probably due to urea permeating the hydrophobic core of the membrane by simple diffusion. The apparent activation energy, EA, of urea transport after maximal inhibition was 59 kJ mol-1, whereas in control cells EA was 34 kJ mol-1 at 1 M and 12 kJ mol-1 at 1 mM urea. In net efflux experiments with no extracellular urea, the permeability coefficient remained constantly high, independent of a variation of intracellular urea between 1 and 500 mM, which indicates that the urea transport system is asymmetric. It is concluded that urea permeability above the ground permeability is due to facilitate diffusion and not to diffusion through nonspecific leak pathways as suggested previously.  相似文献   

13.
Frostbite was produced in the skin of five Hanford Miniature Swine by exposing local areas to chilled air (-75 degrees C) for 1, 3, 5, 10, or 20 min. A copper-constantan thermocouple was inserted into the dermis to measure the temperature. The mean freezing time (the time required to reach 0 degrees C) was approximately 1.9 min. The mean lowest temperatures were 8.8, -15.7, -20.9, -22.5, and -23.4 degrees C for the 1-, 3-, 5-, 10-, and 20-min freezes, respectively. The mean times to rewarm the skin to 25 degrees C were 3.1, 4.5, 5.5, 7.0, and 8.6 min for the 1-, 3-, 5-, 10-, and 20-min freezes, respectively. Significant linear correlations existed between duration of freeze and rewarming times, duration of freeze and lowest temperature, and lowest temperature and rewarming times.  相似文献   

14.
Samples of 1 M KCl solution and 10 samples of intact frog striated muscle were studied at 4-7 degrees C and/or at 21-22 degrees C. Field inhomogeneity was minimized by using small sample volumes and by using a superconducting magnet designed specifically to provide highly homogeneous fields. In the present experiments, magnetic field inhomogeneity was measured to contribute less than 15% to the free induction decay observed for intracellular 39K. The signal-to-noise ratio of the measurements was enhanced by means of extensive time-averaging. The rates of nuclear relaxation for 39K in aqueous solution were 22 +/- 3 (mean +/- 95% confidence limits) s-1 at 4-7 degrees C and 15 +/- 2 s-1 at 21-22 degrees C. For intracellular 39K, (1/T2) was measured to be 327 +/- 22 s-1 and 229 +/- 10 s-1 at the lower and higher temperatures, respectively. The corresponding values for (1/T1) in the same muscle samples were 198 +/- 31 s-1 and 79 +/- 15 s-1 at 4-7 degrees C and at 21-22 degrees C, respectively. These results for 39K are similar to those previously obtained for intracellular 23Na. Since less than 1% of the intracellular 23Na has been estimated to be immobilized, fractional immobilization of intracellular 39K is also likely to be insubstantial.  相似文献   

15.
The transport of 3-O-methylglucose in white fat cells was measured under equilibrium exchange conditions at 3-O-methylglucose concentrations up to 50 mM with a previously described method (Vinten, J., Gliemann, J. and Osterlind, K. (1976) J. Biol. Chem. 251, 794--800). Under these conditions the main part of the transport was inhibitable by cytochalasin B. The inhibition was found to be of competitive type with an inhibition constant of about 2.5 . 10(-7) M, both in the absence and in the presence of insulin (1 micrometer). The cytochalasin B-insensitive part of the 3-O-methylglucose permeability was about 2 . 10(-9) cm . s-1, and was not affected by insulin. As calculated from the maximum transport capacity, the half saturation constant and the volume/surface ratio, the maximum permeability of the fat cell membrane to 3-O-methylglucose at 37 degrees C and in the presence of insulin was 4.3 . 10(-6) cm . s-1. From the temperature dependence of the maximum transport capacity in the interval 18--37 degrees C and in the presence of insulin, an Arrhenius activation energy of 14.8 +/- 0.44 kcal/mol was found. The corresponding value was 13.9 +/- 0.89 in the absence of insulin. The half saturating concentration of 3-O-methylglucose was about 6 mM in the temperature interval used, and it was not affected by insulin, although this hormone increased the maximum transport capacity about ten-fold to 1.7 mmol . s-1 per 1 intracellular water at 37 degrees C.  相似文献   

16.
Human red cell permeability to the homologous series of methanol, ethanol, n-propanol, n-butanol, and n-hexanol was determined in tracer efflux experiments by the continuous flow tube method, whose time resolution is 2-3 ms. Control experiments showed that unstirred layers in the cell suspension were less than 2 X 10(-4) cm, and that permeabilities less than or equal to 10(-2) cm s-1 can be determined with the method. Alcohol permeability varied with the chain length (25 degrees C): Pmeth 3.7 X 10(-3) cm s-1, Peth 2.1 X 10(-3) cm s-1, Pprop 6.5 X 10(-3) cm s-1, Pbut less than or equal to 61 X 10(-3) cm s-1, Phex 8.7 X 10(-3) cm s-1. The permeability for methanol, ethanol, and n- propanol was concentration independent (1-500 mM). The permeability to n-butanol and n-hexanol, however, increased above the upper limit of determination at alcohol concentrations of 100 and 25 mM, respectively. The activation energies for the permeability to methanol, n-propanol, and n-hexanol were similar, 50-63 kJ mol-1. Methanol permeability was not reduced by p-chloromercuribenzene sulfonate (PCMBS), thiourea, or phloretin, which inhibit transport of water or hydrophilic nonelectrolytes. It is concluded (a) that all the alcohols predominantly permeate the membrane lipid bilayer structure; (b) that both the distribution coefficient and the diffusion coefficient of the alcohols within the membrane determine the permeability, and (c) that the relative importance of the two factors varies with changes in the chain length.  相似文献   

17.
The Ca2+-binding kinetics of fura-2 and azo-1 were studied using temperature-jump relaxation methods. In 140 mM KCl at 20 degrees C, the association and dissociation rate constants for fura-2 were 6.02 x 10(8) M-1s-1 and 96.7 s-1, respectively. The fura-2 kinetics were insensitive to pH over the range 7.4 to 8.4. Azo-1 was studied in 140 mM KCl, at pH 7.4, at 10 degrees and 20 degrees C. At 10 degrees C, azo-1 exhibited association and dissociation rate constants of 1.43 x 10(8) M-1s-1 and 777.9 s-1, respectively; while at 20 degrees C, the corresponding values were 3.99 x 10(8) M-1s-1 and 1,177 s-1. The kinetic results demonstrate that fura-2 and azo-1 are well suited to monitoring rapid changes in intracellular [Ca2+].  相似文献   

18.
Interaction of Sindbis virus glycoproteins during morphogenesis.   总被引:14,自引:9,他引:5       下载免费PDF全文
In cells infected with the Sindbis temperature-sensitive mutants ts-23 and ts-10 (complementation group D), which contain a defect in the envelope glycoprotein E1, the precursor polypeptide PE2 is not cleaved to the envelope glycoprotein E2 at the nonpermissive temperature. This defect is phenotypically identical to the defect observed in the complementation group E mutant, ts-20. The lesion in ts-23 is reversible upon shift to permissive temperature, whereas that of ts-10 is not. Antiserum against whole virus, E1, or E2 also prevents the cleavage of PE2 in cells infected with wild-type Sindbis virus. Because the cleavage of PE2 is inhibited by the lesion in mutants that are genotypically distinct and by anti-E1 or -E2 serum, it appears that PE2 and E1 exist as a complex in the membrane of the infected cell.  相似文献   

19.
Fluorescence recovery after photobleaching with an unmodified confocal laser scanning microscope (confocal FRAP) was used to determine the diffusion properties of network forming biological macromolecules such as aggrecan. The technique was validated using fluorescein isothiocyanate (FITC)-labeled dextrans and proteins (molecular mass 4-2000 kDa) at 25 degrees C and with fluorescent microspheres (207 nm diameter) over a temperature range of 5-50 degrees C. Lateral diffusion coefficients (D) were independent of the focus position, and the degree and extent of bleach. The free diffusion coefficient (Do) of FITC-aggrecan determined by confocal FRAP was 4.25 +/- 0.6 x 10(-8) cm2 s-1, which is compatible with dynamic laser light scattering measurements. It appeared to be independent of concentration below 2.0 mg/ml, but at higher concentrations (2-20 mg/ml) the self-diffusion coefficient followed the function D = Do(e)(-Bc). The concentration at which the self-diffusion coefficient began to fall corresponded to the concentration predicted for domain overlap. Multimolecular aggregates of aggrecan ( approximately 30 monomers) had a much lower free diffusion coefficient (Do = 6.6 +/- 1.0 x 10(-9) cm2 s-1) but showed a decrease in mobility with concentration of a form similar to that of the monomer. The method provides a technique for investigating the macromolecular organization in glycan-rich networks at concentrations close to those found physiologically.  相似文献   

20.
The reductant of ferricytochrome c2 in Rhodopseudomonas sphaeroides is a component, Z, which has an equilibrium oxidation-reduction reaction involving two electrons and two protons with a midpoint potential of 155 mV at pH 7. Under energy coupled conditions, the reduction of ferricytochrome c2 by ZH2 is obligatorily coupled to an apparently electrogenic reaction which is monitored by a red shift of the endogeneous carotenoids. Both ferricytochrome c2 reduction and the associated carotenoid bandshift are similarly affected by the concentrations of ZH2 and ferricytochrome c2, pH, temperature the inhibitors diphenylamine and antimycin, and the presence of ubiquinone. The second-order rate constant for ferricytochrome c2 reduction at pH 7.0 and at 24 degrees C was 2 - 10(9) M-1 - s-1, but this varied with pH, being 5.1 - 10(8) M-1 = s-1 at pH 5.2 and 4.3 - 10(9) M-1 - s-1 at pH 9.3. At pH 7 the reaction had an activation energy of 10.3 kcal/mol.  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