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 15 毫秒
1.
The particle weight (molecular weight) of phiNS11 was determined from the sedimentation coefficient, diffusion coefficient, and partial specific volume of the phage. The sedimentation coefficient of the phage (S(0)20, W) is 416 +/- 2.7S. The diffusion coefficient D(0)20, W), which was determined by quasielastic light scattering measurement, is (0.57 +/- 0.03) x 10(-7) cm2/s. The partial specific volume was determined by the mechanical oscillation technique to be 0.747 +/- 0.007 cm3/g. Based on these values, the particle weight of the phage was calculated to be (70.3 +/- 4.3) x 10(6) daltons, which agrees well with the particle weight (69--72 x 10(6) daltons) estimated from the molecular weight of phage DNA and the content of DNA. The Stokes radius of the phage particle was calculated to be 37.7 +/- 2 nm and hydration of the phage was estimated to be 1.18 cm3/g of dry phage. From the particle weight and the chemical composition of the phage, we estimated that one phage particle contains one double-stranded DNA molecule, 16,000 residues of fatty acid, 72 protein I molecules, 920 protein II, 42 protein III, 48 protein IV, 290 protein V molecules, and 3,700 molecules of polyamines.  相似文献   

2.
An image-based technique of fluorescence recovery after photobleaching (video-FRAP) was used to measure the lateral diffusion coefficients of a series of nine fluorescent probes in two model lipid bilayer systems, dimyristoylphosphatidylcholine (DMPC) and DMPC/cholesterol (40 mol%), as well as in human stratum corneum-extracted lipids. The probes were all lipophilic, varied in molecular weight from 223 to 854 Da, and were chosen to characterize the lateral diffusion of small compounds in these bilayer systems. A clear molecular weight dependence of the lateral diffusion coefficients in DMPC bilayers was observed. Values ranged from 6.72 x 10(-8) to 16.2 x 10(-8) cm2/s, with the smaller probes diffusing faster than the larger ones. Measurements in DMPC/cholesterol bilayers, which represent the most thorough characterization of small-solute diffusion in this system, exhibited a similar molecular weight dependence, although the diffusion coefficients were lower, ranging from 1.62 x 10(-8) to 5.60 x 10(-8) cm2/s. Lateral diffusion measurements in stratum corneum-extracted lipids, which represent a novel examination of diffusion in this unique lipid system, also exhibited a molecular weight dependence, with values ranging from 0.306 x 10(-8) to 2.34 x 10(-8) cm2/s. Literature data showed that these strong molecular weight dependencies extend to even smaller compounds than those examined in this study. A two-parameter empirical expression is presented that describes the lateral diffusion coefficient in terms of the solute's molecular weight and captures the size dependence over the range examined. This study illustrates the degree to which small-molecule lateral diffusion in stratum corneum-extracted lipids can be represented by diffusion in DMPC and DMPC/cholesterol bilayer systems, and may lead to a better understanding of small-solute transport across human stratum corneum.  相似文献   

3.
1. A mol.wt. of 40030 +/- 830 has been estimated for phosphoglycerate kinase in concentrations less than 0.1 g/100 cm3 comparing favourably with expected values from X-ray diffraction measurements by 10% lower than the previously reported molecular weights made at higher concentrations. 2. The so20w, was estimated to be 3.12(+/-0.02)x10(-13)s and the coefficient had a low concentration dependency giving a g value (concentration-dependency) of 2.3 +/- 1.6cm3 .g-1. This agrees with previous qualitative observations. 3. By using fluctuation-intensity spectroscopy, the D20,w was estimated to be 7.4(+/-0.2)x10(-11)m2.s-1, and this was indistinguishable from the D20,w calculated from ultracentrifuge results. The water of hydration was estimated to be 0.46 g/g of protein. 4. It is inferred from the estimates that phosphoglycerate kinase associates with an interaction coefficient at 20 degrees C for monomer/dimer of between 10 and 12 cm3.g-1. 5. The ratio of molecular asymmetry (a/b) was estimated to be 2.5+/-0.2 from the values of D20,w and water of hydration. This compares favourably with the ratio from the overall dimensions estimated from X-ray diffraction measurements.  相似文献   

4.
The apparent cytoplasmic proton diffusion coefficient was measured using pH electrodes and samples of cytoplasm extracted from the giant neuron of a marine invertebrate. By suddenly changing the pH at one surface of the sample and recording the relaxation of pH within the sample, an apparent diffusion coefficient of 1.4 +/- 0.5 x 10(-6) cm2/s (N = 7) was measured in the acidic or neutral range of pH (6.0-7.2). This value is approximately 5x lower than the diffusion coefficient of the mobile pH buffers (approximately 8 x 10(-6) cm2/s) and approximately 68x lower than the diffusion coefficient of the hydronium ion (93 x 10(-6) cm2/s). A mobile pH buffer (approximately 15% of the buffering power) and an immobile buffer (approximately 85% of the buffering power) could quantitatively account for the results at acidic or neutral pH. At alkaline pH (8.2-8.6), the apparent proton diffusion coefficient increased to 4.1 +/- 0.8 x 10(-6) cm2/s (N = 7). This larger diffusion coefficient at alkaline pH could be explained quantitatively by the enhanced buffering power of the mobile amino acids. Under the conditions of these experiments, it is unlikely that hydroxide movement influences the apparent hydrogen ion diffusion coefficient.  相似文献   

5.
The purification and properties of isocitrate lyase from Chlorella   总被引:11,自引:5,他引:6  
1. Isocitrate lyase (threo-d(s)-isocitrate glyoxylate-lyase, EC 4.1.3.1) has been purified from acetate-adapted cells of Chlorella pyrenoidosa. 2. The final preparation was homogeneous by the criteria of sedimentation, diffusion and polyacrylamide-gel electrophoresis. 3. The sedimentation coefficient (S(20,w)) was 9.04x10(-13)sec. and the diffusion coefficient (D(20,w)) 4.62x10(-7)cm.(2)/sec.; from these values the molecular weight of the enzyme was calculated to be 170000 and its Stokes radius to be 4.63x10(-7)cm. 4. The elution of the enzyme from Sephadex G-100 was studied and estimates of molecular weight and Stokes radius were obtained from the elution data. 5. The turnover number of the enzyme was 5950moles of glyoxylate formed/min./mole of enzyme at 30 degrees . 6. With threo-d(s)(+)-isocitrate as substrate, the K(m) of the enzyme was 0.023mm.  相似文献   

6.
采用时间序列分析法,对江西省修水县退化石灰岩红壤区4种主要重建森林类型的土壤(0~40 cm)含水量与降水量、蒸发量之间的相关关系进行了研究.结果表明:研究区各重建森林前一个月的土壤含水量对当月整个土壤剖面(0~40 cm)平均土壤含水量和当月20~40 cm土层土壤含水量具有较强影响,对当月0~10 cm土层土壤含水量的影响较小;研究区各重建森林20~40 cm土层土壤含水量主要受当月降水量的影响,0~10 cm土层土壤含水量主要受当月蒸发量的影响;在4种森林重建模式中,枫香纯林当月降水量与当月土壤含水量的相关系数最大,混交林模式土壤含水量的自相关系数大于纯林模式.  相似文献   

7.
D Revie  M E Dahmus 《Biochemistry》1979,18(9):1813-1820
A heat-stable protein (HSF) that stimulates the activity of lamb thymus RNA polymerase II has been purified 2500-fold and partially characterized. This factor stimulates the activity of RNA polymerase II up to 13 times and retains complete activity when heated at 90 degrees C for 5 min. Stimulation is observed only in the presence of RNA polymerase II and requires native DNA as template. The stimulatory factor has a sedimentation coefficient of 2.7 S, a diffusion coefficient of 9.55 x 10(-7) cm2/s, and an isoelectric point of 8.0. Calculated from the sedimentation and diffusion data, the factor has a molecular weight of about 24,000. Electrophoresis of the purified factor on polyacrylamide gels in the presence of sodium dodecyl sulfate results in a single band corresponding to a molecular weight of 25,000. The number-average length of the RNA synthesized by RNA polymerase II is increased in the presence of the factor. Sedimentation velocity and exclusion chromatography experiments suggest that the stimulatory factor interacts with RNA polymerase II. These results suggest that the factor stimulates RNA synthesis through a direct interaction with RNA polymerase II. The stoichiometry of the HSF-RNA polymerase binding appears to be about 1:1. HSF is located in the nucleus, as determined by cell fractionation studies.  相似文献   

8.
The nitric oxide synthase inhibitor N(G)-nitro-L-arginine methyl ester (l-NAME) increased vascular resistance (VR) 10% above baseline of 3.08+/-0.08 (n=11) mmHg/mL/min at 10 mg/kg and 20% above 3.05+/-0.08 (n=9) at 50 mg/kg in anesthetized toads (Bufo marinus). Blood pressure was unaffected by either dose of L-NAME. Blood flow decreased at the higher dose of L-NAME. L-arginine (300 mg/kg) reversed the effects of L-NAME on VR and blood flow in toads treated with 10 mg/kg but not with 50 mg/kg. Injection of 50 mg/kg L-NAME into empty-bladder toads produced a 10% decrease in water uptake, J(v), resulting in a J(v) of 1,267+/-11 cm(3)/cm(2)/s x 10(-7) (n=9) compared to 1,385+/-12 (n=8) for controls. Injection of 10 microg/kg angiotensin II (ANG II) increased J(v) 15% across the pelvic patch (J(v), cm(3)/cm(2)/s x 10(-7)), resulting in a J(v) of 1,723+/-12 cm(3)/cm(2)/s x 10(-7) (n=8) compared to 1,471+/-12 (n=8) for controls. It is hypothesized that during cutaneous drinking blood flow into the capillary bed of the pelvic patch is regulated by nitric oxide and ANG II.  相似文献   

9.
The lateral mobility of the lipid analog N-4-nitrobenzo-2-oxa-1,3 diazole phosphatidylethanolamine and of the integral protein glycophorin in giant dimyristoylphosphatidylcholine vesicles was studied by the photobleaching technique. Above the temperature of the chain-melting transition (Tm = 23 degrees C), the diffusion coefficient, Dp, of the protein [Dp = (4 +/- 2) X 10(-8) cm2/s at 30 degrees C] was within the experimental errors equal to the corresponding values DL of the lipid analog. In the P beta 1 phase the diffusion of lipid and glycophorin was studied as a function of the probe and the protein concentration. (a) At low lipid-probe content (cL less than 5 mmol/mol of total lipid), approximately 20% of the probe diffuses fast (D approximately equal to 10(-8) - 10(-9) cm2/s), while the mobility of the rest is strongly reduced (D less than 10(-10) cm2/s). At a higher concentration (cp approximately 20 mmol), all probe is immobilized (D less than 10(-10) cm2/s). (b) Incorporation of glycophorin up to cp = 0.4 mmol/mol of total lipid leads to a gradual increase of the fraction of mobile lipid probe due to the lateral-phase separation into a pure P beta 1 phase and a fraction of lipid that is fluidized by strong hydrophilic lipid-protein interaction. (c) The diffusion of the glycophorin molecules is characterized by a slow and a fast fraction. The latter increases with increasing protein content, which is again due to the lateral-phase separation caused by the hydrophilic lipid-protein interaction. The results are interpreted in terms of a fast transport along linear defects in the P beta 1 phase, which form quasi-fluid paths for a nearly one dimensional and thus very effective transport. Evidence for this interpretation of the diffusion measurements is provided by freeze-fracture electron microscopy.  相似文献   

10.
We performed single perfusion experiments in the small intestine of rats in order to prove that the two-dimensional laminar flow model is suitable to determine the membrane permeability coefficient and the reflection coefficient. We used progesterone as an aqueous-diffusion-limited drug, urea as a membrane transport-limited drug and the tritiated water as an intermediate substance. The membrane permeability coefficient for progesterone was calculated to be 3.6 X 10(-4) cm/s. This value did not change when the thickness of the aqueous diffusion layer was altered by increasing the perfusion rate 10-fold. It was directly demonstrated that the two-dimensional laminar flow model was suitable to analyze the data of intestinal perfusion experiments. Membrane permeability coefficients for urea and tritiated water were determined to be 3.4 X 10(-5) cm/s and 8.9 X 10(-5) cm/s, respectively. In the presence of water absorption with the hypotonic perfusion solution, the reflection coefficient for urea was 0.84. This value is thought to be theoretically reasonable, suggesting the usefullness of the two-dimensional laminar flow model to obtain the reflection coefficient in the intestinal membrane.  相似文献   

11.
R Bülow  P Overath  J Davoust 《Biochemistry》1988,27(7):2384-2388
The membrane form of the variant surface glycoprotein (mfVSG) is anchored in the plasma membrane of Trypanosoma brucei by a dimyristoylphosphatidylinositol residue connected via a glycan to the COOH-terminal amino acid. The glycoprotein molecules are tightly packed, forming a coat that is impenetrable to lytic serum components. Lateral diffusion of mfVSG was measured by the fluorescence recovery after photobleaching technique. mfVSG labeled on the cell surface with rhodamine-conjugated anti-VSG Fab fragments showed a diffusion coefficient of 1 X 10(-10) cm2/s at 37 degrees C and of 0.7 X 10(-10) cm2/s at 27 degrees C. About 80% of the molecules were mobile. Affinity-purified mfVSG molecules implanted into the plasma membrane of baby hamster kidney cells exhibited a similar mobility to that found in the trypanosome coat [D = (0.4-0.7) X 10(-10) cm2/s at 4 degrees C]. Phospholipid mobility in the plasma membrane of trypanosomes was characterized by a diffusion coefficient of 2.2 X 10(-9) cm2/s at 37 degrees C. It is concluded that mfVSG mobility in the surface coat of the parasite is rapid and comparable to that of other membrane-bound glycoproteins but slower than that of phospholipids.  相似文献   

12.
Pulmonary emphysema is currently thought to be due to an elastase-antielastase imbalance with resultant destruction of alveolar structures. The present study was aimed at testing whether alpha 1-proteinase inhibitor (alpha 1 PI) is the major component of the antielastase screen of the lower respiratory tract of healthy subjects. Bronchoalveolar lavage was performed in 8 nonsmokers (27.8 +/- 3.8 years) and 9 smokers (25 +/- 0.96 years). The lavage fluids were tested for leukocyte and pancreatic elastase inhibitory capacity (LEIC and PEIC) and immunoreactive alpha 1 PI and bronchial inhibitor (brI) content. The mean +/- s.e.m. levels of LEIC, PEIC, alpha 1 PI and brI were 0.16 +/- 0.039, 0.042 +/- 0.006, 0.09 +/- 0.007 and 0.013 +/- 0.002 mol/mol albumin, respectively. Thus, on the average, the molar concentration of brI was about 14% that of alpha 1 PI. The difference between LEIC and alpha 1 PI did not reach statistical significance (P = 0.0503). The PEIC was however significantly lower than the alpha 1 PI levels (P less than 0.05), indicating that the lavage fluids contained both active and inactive alpha 1 PI. Nonsmokers and smokers did not differ in their LEIC, PEIC, alpha 1 PI and brI levels. When the data were examined on an individual basis, the subjects could be divided into 2 groups: group I (n = 9; 3 nonsmokers, 6 smokers) whose LEIC/alpha 1 PI molar ratios were higher than unity and group II (n = 8; 5 nonsmokers, 3 smokers) whose LEIC/alpha 1 PI molar ratios were equal or lower than unity. Group I subjects had significantly higher LEIC values (0.26 +/- 0.05 mol elastase inhibited/mol albumin) than group II individuals (0.055 +/- 0.006; P less than 0.001) but the two groups had similar levels of immunoreactive alpha 1 PI (0.09 and 0.08 mol alpha 1 PI/mol albumin for group I and II, respectively), functionally active alpha 1 PI (percentage of active alpha 1 PI: 53% and 37% for group I and II, respectively) and immunoreactive brI (0.016 and 0.010 mol brI/mol albumin for group I and II, respectively). These results suggested that the lavage fluids from group I contained significant amounts of undefined leukocyte elastase inhibitor(s). Gel filtration of a lavage fluid from group I showed that the undefined elastase inhibitor(s) co-eluted with brI. Most of the lavage fluids were still able to inhibit leukocyte elastase following removal of alpha 1 PI by perchloric acid precipitation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

13.
Diffusion of ionic and nonionic species in multilayered tissues plays an important role in the metabolic processes that take place in these tissues. To create a mathematical model of these diffusion processes, we have chosen as an example hydrogen-bicarbonate ion pair diffusion within the mammalian cornea. This choice was based on the availability of experimental data on this system. The diffusion coefficient of the hydrogen-bicarbonate ion pair in corneal stroma and epithelium is calculated from the observed change in pH in the stroma when conditions at the corneal anterior epithelial surface are changed while the posterior surface is continually bathed with a Ringer's solution in equilibrium with a CO2-gas air mixture. Matching experimental results to a mathematical model of the cornea as a two-layer diffusion system yields, at 37 degrees C, a diffusion coefficient of the hydrogen-bicarbonate ion pair of 2.5 x 10(-6) cm2/s in the stroma and 0.4 x 10(-6) cm2/s in the epithelium. Application of the Nernst-Einstein equation to these data gives the following diffusion coefficients in the two layers: 1) stroma, D(H+) = 11.8 x 10(-6) cm2/s; D(HCO3-) = 1.5 x 10(-6) cm2/s; and 2) epithelium, D(H+) = 1.9 x 10(-6) cm2/s; D(HCO3-) = 0.22 x 10(-6) cm2/s.  相似文献   

14.
Bovine myelin basic protein has been investigated with regard to its solution behavior, circular dichroism and 220 MHz PMR spectral properties. At pH 4.8 gamma/2=0.1 acetate buffer, light scattering yielded a Mr of 17 700 and a virial coefficient of 1.0-10(-4) mol-ml/g2. Above pH 7.0 the protein was found to aggregate to higher mol. wt species. Sedimentation experiments at pH 4.8 yielded s degrees 20,w of 1.27 S at gamma/2=0.1 and 1.46 S at gamma/2=0.35. The diffusion coefficient determined from ultracentrifugal experiments was 7.25-10(-7) cm2/s at gamma/2=0.1 and 0.35. The value of f/f0 from diffusion at pH 4.8 and gamma/2=0.35 was 1.64, corresponding to an axial ratio of 11 to 1. The radius of gyration was calculated as 4.28 nm and the root mean square end to end distance was 10.5 nm. At pH 9.0, gamma/2=0.1, s degrees 20,w was 1.71 S and D degrees 20,w was estimated at 7.4-10(-7) cm2/s. The behavior at pH 9.0 reverted to the behavior at pH 4.8 when the pH was readjusted. The E1%/1cm=5.64 at 276.4 nm and 225 at 196 nm. Titration of the protein with trifluoroethanol elicited three distinct regions of conformation stability having increasing helical content as the mol fraction of trifluoroethanol increased. The results of the present study have permitted some comparison of analogous properties and conformational behavior with the basic membrane protein cytochrome c.  相似文献   

15.
R Peters 《The EMBO journal》1984,3(8):1831-1836
Fluorescence microphotolysis was used to measure nucleocytoplasmic flux in single rat hepatocytes for a series of dextrans ranging in molecular mass from 3 to 150 kd. The cytoplasmic translational diffusion coefficient DC and the nucleoplasmic diffusion coefficient DN of a 62-kd dextran were also determined. DC was approximately 2 X 10(-8) and DN approximately 3 X 10(-8) cm2/s, i.e., 1/20-1/15 of the value in free solution. The mobile fraction amounted to 0.7-0.8 in measurements of both intracellular diffusion and nucleo-cytoplasmic flux. The flux of dextrans from cytoplasm to nucleus depended inversely on molecular mass with an exclusion limit between 17 and 41 kd suggesting that the nuclear envelope has functions of a molecular sieve. Employing the Pappenheimer-Renkin equations, a functional pore radius of 50-56 A was derived. By comparison with recent measurements on isolated liver cell nuclei, large quantitative differences between the intracellularly located and the isolated nucleus were revealed.  相似文献   

16.
17.
Dihydroorotase (DHOase) catalyzes the third step in eukaryotic de novo pyrimidine biosynthesis. In mammalian cells, this enzyme activity is carried by a large chimeric protein, CAD, that also catalyzes the first two steps in the pathway: glutamine-dependent carbamyl phosphate synthetase (CPSase) and aspartate transcarbamylase (ATCase). Controlled elastase cleavage of CAD released a 44,000 +/- 2,000-dalton proteolytic fragment which catalyzed only the dihydroorotase reaction. We have devised a rapid and simple method for the isolation of the DHO domain from elastase digests. The domain, which was obtained in 36% yield, was found to be homogeneous by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and isoelectric focusing. The domain was also characterized by amino acid analysis and analytical high pressure liquid chromatography peptide mapping. The amino terminus of both the DHO domain and intact CAD was blocked suggesting that this domain is located at the extreme amino terminus of the CAD polypeptide, a result consistent with the suspected juxtaposition of domains as DHO-CPS-ATC. The isoelectric point of the DHO domain was 5.1, while that of the ATC domain was 9.4, so that the ends of the CAD polypeptide are oppositely charged at physiological pH. Immunoblotting with DHO domain-specific antibodies showed that a 47-kDa species was generated in the early stages of controlled proteolysis of CAD. Thus there are two elastase cleavage sites within a 3-kDa connecting region that links the DHO and CPS domains. The domain was shown by atomic absorption spectrophotometry and by isolating a 65Zn-containing DHO domain from mammalian cells grown in the presence of the radionuclide to contain 1 g eq of tightly bound zinc in each polypeptide chain. Zinc was not found in any other CAD domain. Chelating agents inhibit dihydroorotase activity of the isolated domain supporting the conclusion, based on studies of intact CAD by others, that zinc participates in catalysis. At moderate protein concentrations the DHO domain was a 88,000 dimer with a Stokes radius of 37.6 A, a S20,w = 5.1 X 10(-13) s, a diffusion coefficient of 3.17 X 10(-7) cm2 s-1, and a frictional ratio of 1.26. On dilution the dimer dissociated and was in rapid concentration-dependent equilibrium with a 43,500 monomer. The hydrodynamic parameters of the monomer have also been estimated (Stokes radius of 29.8 A, D20,w = 4.11 X 10(-7) cm2 s-1, and f/f0 1.21).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
An amphiphilic hydrogel of poly(2-hydroxyethyl methacrylate) cross-linked with tetraethyleneglycol diacrylate (TEGDA) was synthesized to contain the hydrophobic monomer 3-(trimethoxy-silyl) propyl methacrylate (PMA) and the pH-responsive, hydrophilic monomer N',N'-dimethylaminoethyl methacrylate (DMAEMA). The gels were separately loaded with two biomolecular probes, insulin and protamine, via both physical entrapment and equilibrium imbibition methods. The release profiles for these biomolecular probes, possessing similar MW (5.7 and 4-6 kDa, respectively) but different pI's (5.3 and 10.0, respectively), were investigated with respect to variation in the pH of the bathing medium as well as the DMAEMA content, and the cross-link density of the hydrogel. Gels exhibited classical Fickian diffusion release profiles. For a typical gel composition 66:15:10:09 mol % (HEMA:DMAEMA:PMA:TEGDA), as the pH of the release media decreased from 7.3 to 4.0, the rate of release of both biomolecular probes increased. When loaded via entrapment, the insulin release rate increased ca. 4-fold (1.0-3.7 x 10(-7) cm(2) s(-1)), whereas that of protamine increased 10-fold (0.3-3.3 x 10(-7) cm(2) s(-1)). When loaded by imbibition, the insulin diffusion coefficient increased 2-fold (3.8-7.2 x 10(-7) cm(2) s(-1)), whereas that of protamine increased 3-fold (1.9-5.5 x 10(-7) cm(2) s(-1)). The reduction of pH, through its protonation of the gel network, has a more dramatic influence on protamine release, the result of its higher pI (10.0) compared to that of insulin (5.3). As the DMAEMA content of the hydrogel was increased from 0 to 20 mol %, the diffusion coefficient of protamine increased by ca. 7-fold (1.7-12.2 x 10(-7) cm(2) s(-1)), whereas that of insulin increased only ca. 2-fold (1.7-4.0 x 10(-7) cm(2) s(-1)). This differential release confirms the role of internal protonation in effecting the greater release of the protonated drug molecule. Increasing the TEGDA content from 3 to 15 mol % reduced the diffusion coefficient ca. 3-fold for insulin (1.6-0.5 x 10(-7) cm(2) s(-1)) and 5-fold for protamine (4.0-0.8 x 10(-7) cm(2) s(-1)). The final D(ip) at 15 mol % TEGDA suggests that the smaller mesh size offsets any differential release that arises from protonation. The presence of PMA in the hydrogel formulation, which contributes additional cross-links by reason of the formation of siloxane macromers, did not change the usually observed Fickian diffusion mechanism.  相似文献   

19.
Motion of cytochrome c bound to giant (2-10-micron diam) mitochondria isolated from the waterbug Lethocerus indicus was examined using the technique of fluorescence recovery after photobleaching. Fluorescent cytochrome c was exchanged for native cytochrome c through partly damaged outer membrane. Recovery profiles were not statistically different when the fluorescence from iron-free cytochrome c or fluorescein-labeled cytochrome c was used and were essentially the same in the presence or absence of an uncoupler. In the presence of excess porphyrin cytochrome c, the apparent diffusion coefficient was 6 X 10(-11) cm2/s in 0.3 M sucrose-mannitol-EDTA and 3 X 10(-10) cm2/s in 0.10 M KCl/0.10 M sucrose. At concentrations of porphyrin cytochrome c that are stoichiometric with cytochrome c oxidase and for mitochondria in which excess cytochrome c was washed away, two components were observed in the recovery profile. The diffusion coefficient of the fast component was 1 X 10(-10) cm2/s. The second component showed no recovery during the time scale of measurement (D less than 10(-12) cm2/s). We speculate on the origin of the immobile fraction.  相似文献   

20.
Structural failure of the erythrocyte membrane in shear deformation occurs when the maximum shear resultant (force/length) exceeds a critical value, the yield shear resultant. When the yield shear resultant is exceeded, the membrane flows with a rate of deformation characterized by the plastic viscosity coefficient. The temperature dependence of the yield shear resultant and the plastic viscosity coefficient have been measured over the temperature range 10-40 degrees C. Over this range the yield shear resultant does not change significantly (+/- 15%), but the plastic viscosity coefficient changes exponentially from a value of 1.3 X 10(-2) surface poise (dyn s/cm) at 10 degrees C to a value of 6.2 X 10(-4) surface poise (SP) at 40 degrees C. The different temperature dependence of these two parameters is not surprising, inasmuch as they characterize different molecular events. The yield shear resultant depends on the number and strength of intermolecular connections within the membrane skeleton, whereas the plastic viscosity depends on the frictional interactions between molecular segments as they move past one another in the flowing surface. From the temperature dependence of the plastic viscosity, a temperature-viscosity coefficient, E, can be calculated: eta p = constant X exp(--E/RT). This quantity (E) is related to the probability that a molecular segment can "jump" to its next location in the flowing network. The temperature-viscosity coefficient for erythrocyte membrane above the elastic limit is calculated to be 18 kcal/mol, which is similar to coefficients for other polymeric materials.  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