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 kinetic and concentration dependences of erythrocyte vesiculation and hemolysis induced by sodium dodecyl sulfate were studied. The similarity of the slopes of the dose dependence of the SDS-induced vesiculation and slow hemolysis rates in the double logarithmic coordinates suggested a close relation between the processes of vesiculation and pore formation for slow hemolysis by the detergent. Further evidence of the competitive nature of the detergent-induced vesiculation and fast hemolysis by sodium dodecyl sulfate was obtained. The phenomenon of partial hemolysis proceeding at a rate comparable to that of cell vesiculation is explained in terms of the competition between hemolysis and vesiculation, without resorting to erythrocyte heterogeneity. New vesicular-competitive hemolysis is described. Based on it, the action of different hemolysis-inducing agents is analysed.  相似文献   

2.
The thermodynamic parameters for the interaction of the anionic detergent sodium n-dodecyl sulphate (SDS) with H2B at pH 3.2, 6.4 and 10 have been measured at 27 degrees C and 37 degrees C by equilibrium dialysis to determine the Gibbs energies of detergent binding. The data have been used to obtain the enthalpy of interaction from the temperature dependence of the equilibrium constants from the Van't Hoff relation. The enthalpy of interaction between H2B and SDS is endothermic at pH 3.2, 6.4 and 10. The shapes of the enthalpy curves at pH 3.2 and 10 show some small exothermic contribution which probably indicates folding of H2B. The interactions of H2B-SDS are dominated by the increase in entropy on detergent binding. The larger negative free energy, enthalpy and entropy changes at pH 6.4 are consistent with greater denaturation relative to pH 3.2 and 10.  相似文献   

3.
The range of pH values for the blood of grasshoppers and of houseflies is 7.2 to 7.6. The range of values for roaches is 7.5 to 8.0. The range for Malacosoma americanum is 6.4 to 7.4; and the range for Bombyx mori is 6.4 to 7.2. From the work of other investigators and from the writer''s results, it is apparent that the pH of insect blood, in general, may vary between 6.4 and 8.0. In the forms observed no correlation exists between blood pH and age, nor between pH and metamorphosis.  相似文献   

4.
The absorbance change at 505 nm was used to monitor the kinetics of violaxanthin deepoxidation in isolated pea (Pisum sativum) chloroplasts under dark conditions at various pH values. In long-term measurements (65 min) a fast and a slow exponential component of the 505-nm absorbance change could be resolved. The fast rate constant was up to 10 times higher than the slow rate constant. The asymptote value of the fast kinetic component was twice that of the slow component. The pH dependency of the parameters of the fast kinetic component was analyzed from pH 5.2 to pH 7.0. It was found that the asymptote value dropped slightly with increasing pH. The rate constant was zero at pH values greater than 6.3 and showed maximum values at pH values less than 5.8. Hill plot analysis revealed a strong positive cooperativity for the pH dependency of the fast rate constant (Hill coefficient nH = 5.3). The results are discussed with respect to published activity curves of violaxanthin deepoxidation.  相似文献   

5.
A continuous spectrophotometric method for monitoring phospholipase D-catalyzed hydrolysis of long acyl chain phosphatidylcholines has been formulated at pH 8.0 in a mixed detergent system using the coupling enzymes choline oxidase and peroxidase. Standard curves for phosphatidylcholine determination in both end-point and rate modes are presented and applied to the estimation of that phospholipid in a solubilized human erythrocyte membrane sample. In rate mode the method is suitable for kinetic study of phospholipase D with physiological substrates in micellar form.  相似文献   

6.
I Vass  S Styring 《Biochemistry》1991,30(3):830-839
The effect of protonation events on the charge equilibrium between tyrosine-D and the water-oxidizing complex in photosystem II has been studied by time-resolved measurements of the EPR signal IIslow at room temperature. The flash-induced oxidation of YD by the water-oxidizing complex in the S2 state is a monophasic process above pH 6.5 and biphasic at lower pHs, showing a slow and a fast phase. The half-time of the slow phase increases from about 1 s at pH 8.0 to about 20 s at pH 5.0, whereas the half-time of the fast phase is pH independent (0.4-1 s). The dark reduction of YD+ was followed by measuring the decay of signal IIslow at room temperature. YD+ decays in a biphasic way on the tens of minutes to hours time scale. The minutes phase is due to the electron transfer to YD+ from the S0 state of the water-oxidizing complex. The half-time of this process increases from about 5 min at pH 8.0 to 40 min at pH 4.5. The hours phase of YD+ has a constant half-time of about 500 min between pH 4.7 and 7.2, which abruptly decreases above pH 7.2 and below pH 4.7. This phase reflects the reduction of YD+ either from the medium or by an unidentified redox component of PSII in those centers that are in the S1 state. The titration curve of the half-times for the oxidation of YD reveals a proton binding with a pK around 7.3-7.5 that retards the electron transfer from YD to the water-oxidizing complex. We propose that this monoprotic event reflects the protonation of an amino acid residue, probably histidine-190 on the D2 protein, to which YD is hydrogen bonded. The titration curves for the oxidation of YD and for the reduction of YD+ show a second proton binding with pK approximately 5.8-6.0 that accelerates the electron transfer from YD to the water-oxidizing complex and retards the process in the opposite direction. This protonation most probably affects the water-oxidizing complex. From the measured kinetic parameters, the lowest limits for the equilibrium constants between the S0YD+ and the S1YD as well as between the S1YD+ and S2YD states were estimated to be 5 and 750-1000, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
Dense Cultures of Neisseria gonorrhoeae in Liquid Medium   总被引:12,自引:2,他引:10       下载免费PDF全文
Cultivation of Neisseria gonorrhoeae was effected in a conical glass culture vessel surrounded by a constant-temperature water jacket, and with facilities for stirring, aeration, and pH measurement and control. With the use of an aerated peptone-based medium, containing polypropylene glycol to prevent foam build-up, the yields obtained over the pH range from 5.8 to 7.4 were determined. The greatest yield was obtained at pH 6.4 when the dry weight was 1.5 g/liter. At pH 7.2 to 7.6, lysis was extensive.  相似文献   

8.
The precipitating of effect of sodium dodecyl sulphate (SDS) on the egg white proteins ovalbumin, conalbumin and lysozyme was studied at 25 degrees C and at different pH values. The proteins precipitated below their respective isolectric points, provided the detergent to protein ratio was appropriate. The pH profile of precipitation was different for the three proteins reflecting net charge differences. The binding of SDS to the proteins was studied with [35S]-labelled SDS and for ovalbumin a ratio of 21--28 SDS molecules/protein molecule, dependent on the concentration of SDS initially used, seem to be required for precipitation at pH 4.5. This number, however, is dependent on pH and increases with an increased positive net charge of the protein. The precipitating effect of SDS was identical for ovalbumin, conalbumin and lysozyme when compared on a gram to gram basis (0.1--0.15 g SDS/g precipitated protein). The precipitated protein was denatured as measured by differential scanning calorimetry, but was also completely redissolved if pH was increased to above the isoelectric point. The precipitating effecto f SDS was also examined at elevated temperatures. The two-phase systems of the proteins induced by SDS at 25 degrees C were heated from 25 degrees C to 90 degrees C at a rate of 1.25 degrees C/min. The precipitation behaviour was similar for the three proteins upon heating. When the SDS concentration was increased the precipitation curves were transferred towards lower temperatures and the courses of precipitation became less sharp. The synergistic effect of SDS and heat on protein precipitation was differentiated by denaturation measurements and radioactive labelling. The ratio SDS to precipitated protein gradually diminished towards higher temperatures but no purely thermal precipitation was found.  相似文献   

9.
Two isoenzymes of ATP:D-fructose-6-phosphate 1-phosphotransferase(phosphofructokinase) are present in germinating cucumber seeds,one in the plastids and the other in the cytosol. Both isoenzymeswere purified and some of their kinetic properties studied.These two isoenzymes differ kinetically, the pH optimum of thecytosolic isoenzyme being 7.2 and that of the plastid isoenzymebeing 8.0. Both isoenzymes are activated by phosphate althoughthe concentration required for activation is much lower forthe plastid isoenzyme than cytosolic isoenzyme. Phosphate increasesthe affinity of the isoenzymes for fructose-6-phosphate andalso changes the sigmoidal kinetics of the plastid isoenzymefor this substrate to hyperbolic kinetics at pH 7.2. The fructose-6-phosphatesaturation kinetics of the cytosolic isoenzyme becomes moresigmoidal with an increase in pH while the opposite is truefor the plastid isoenzyme. The cytosolic isoenzyme has a higheraffinity for fructose-6-phosphate at pH 7.2 than pH 8.0 whilethe affinity of the plastid isoenzyme for fructose-6-phosphateis highest at pH 8.0. Both isoenzymes are inhibited by ATP andthe extent of inhibition is pH dependent. The cytosolic isoenzymeis more sensitive to ATP inhibition at pH 8.0 than pH 7.2 whilethe opposite holds for the plastid isoenzyme. Magnesium alleviatesthe ATP inhibition of the plastid isoenzyme suggesting thatfree ATP is the inhibitory form. In contrast the ATP inhibitionof the cytosolic isoenzyme apparently appears to be caused bythe magnesium-ATP complex. (Received May 19, 1987; Accepted January 18, 1988)  相似文献   

10.
D R McCaslin  C Tanford 《Biochemistry》1981,20(18):5207-5212
When detergent-solubilized proteins interact with hydrophobic or amphiphilic molecules in the presence of detergent micelles, the solubility of the latter species in the micelles must be included in both thermodynamic and kinetic treatments. In this paper, we derive equations which describe the distribution of species present at equilibrium for a system in which a detergent-solubilized protein binds a hydrophobic (or amphiphilic) ligand. We have applied the formalism developed in this paper to the reaction describing the formation of rhodopsin from its apoprotein and 11-cis-retinal. Qualitatively, the results demonstrate that a significant portion of the observed decrease in the extent of recombination for rhodopsin solubilized in either sodium cholate or Tween 80 may be attributed to the partition of retinal into detergent micelles and that a detergent-induced protein denaturation need not be invoked to explain the data. We also discuss results for rhodopsin solubilized in a nonionic detergent (octaethylene glycol n-dodecyl ether) in which the detergent is clearly causing irreversible loss of the capability to recombine with 11-cis-retinal.  相似文献   

11.
The enthalpies of interaction between bovine catalase and sodium n-dodecyl sulphate (SDS) in aqueous solutions of pH 3.2,6.4 and 10.0 have been measured over a range of SDS concentrations by microcalorimetry at 25°C. The enthalpies increase with decreasing pH and with increasing SDS concentration and largely arise from the interations between the anionic head group of SDS and the cactionic amino acid residues on the protein. Chemically modified catalase in which a proportion of carboxylic acid groups have been coupled with either glycine methyl ester or ethylenediamine have been prepared and characterized in terms of their enzymic activities, spectral properties and sedimentation behaviour. The enthalpies of interaction of these catalases with SDS have been studied at pH 6.4. The results of the experiments suggest that the enthalpies of interaction with SDS can be correlated with the ratio of cationic to anionic amino acid residues on the surface of the catalase molecules and hence the nominal net surface charge. The variation in the enthalpy of interaction of catalases with surface charge, as a consequence of variation in pH, differs from the variation with charge at constant pH possibly due to the thermal effect of proton binding to the catalase—complexes.  相似文献   

12.
E Gross  U Hopfer 《Biophysical journal》1999,76(6):3066-3075
The effects of pH on cotransporter kinetics were studied in renal proximal tubule cells. Cells were grown to confluence on permeable support, mounted in an Ussing-type chamber, and permeabilized apically to small monovalent ions with amphotericin B. The steady-state, dinitrostilbene-disulfonate-sensitive current (DeltaI) was Na+ and HCO3- dependent and therefore was taken as flux through the cotransporter. When the pH of the perfusing solution was changed between 6.0 and 8.0, the conductance attributable to the cotransporter showed a maximum between pH 7.25 and pH 7.50. A similar profile was observed in the presence of a pH gradient when the pH of the apical solutions was varied between 7.0 and 8.0 (basal pH lower by 1), but not when the pH of the basal solution was varied between 7.0 and 8.0 (apical pH lower by 1 unit). To delineate the kinetic basis for these observations, DeltaI-voltage curves were obtained as a function of Na+ and HCO3- concentrations and analyzed on the basis of a kinetic cotransporter model. Increases in pH from 7.0 to 8.0 decreased the binding constants for the intracellular and extracellular substrates by a factor of 2. Furthermore, the electrical parameters that describe the interaction strength between the electric field and substrate binding or charge on the unloaded transporter increased by four- to fivefold. These data can be explained by a channel-like structure of the cotransporter, whose configuration is modified by intracellular pH such that, with increasing pH, binding of substrate to the carrier is sterically hindered but electrically facilitated.  相似文献   

13.
Spontaneous vesiculation of aqueous lipid dispersions   总被引:3,自引:0,他引:3  
H Hauser  N Gains  H J Eibl  M Müller  E Wehrli 《Biochemistry》1986,25(8):2126-2134
The swelling properties of lipid mixtures consisting of phosphatidylcholine and a charged single-chain detergent have been studied. The work presented here is confined to lipid mixtures forming smectic lamellar phases in H2O. These mixtures exhibit continuous swelling with increasing water content, provided the surface charge density exceeds a threshold value of about 1-2 microC/cm2. In excess H2O, such mixtures undergo spontaneous vesiculation: unilamellar vesicles form spontaneously when excess H2O or salt solutions of moderate ionic strength (I less than 0.2) are added to the dried film of such lipid mixtures. The resulting dispersion of unilamellar vesicles is usually polydisperse. Its average size depends on the detergent/phospholipid mole ratio, decreasing with increasing detergent content. It is shown that in the phase diagram of three-component systems consisting of phosphatidylcholine, a charged single-chain detergent, and excess H2O there is a compositional range, though narrow, within which the small unilamellar vesicle (diameter less than 100 nm) is the thermodynamically most stable structure. This behavior is characteristic of charged, single-chain detergents of 14 and more C atoms. Many pharmacologically active compounds are amphiphilic and surface-active, and as such, they will orient at phospholipid-water interfaces, imparting a net surface charge to neutral lipid surfaces. It is shown that such drugs exhibit detergent-like action. Mixed films of phosphatidylcholine and a pharmacologically active compound behave similarly to phosphatidylcholine-detergent mixtures: they undergo spontaneous vesiculation when excess H2O or salt solutions of moderate ionic strength are added. In this case, the drug itself induces vesiculation; possible pharmacological implications of this finding are discussed.  相似文献   

14.
The effects of potassium sorbate, sodium hypophosphite, sodium tripolyphosphate, sodium nitrite, and linoleic acid on the germination and outgrowth of Clostridium botulinum type E spores were studied in microcultures. At pH 5.8 to 6.0 in liver veal agar, the germination rate was decreased to nearly zero with 1.0, 1.5, or 2.0% sorbate. At pH 7.0 t 7.2, these levels of sorbate afforded germination and outgrowth of abnormally shaped cells that were defective in cell division. At the high pH range, 0.5 or 1.0% hypophosphite had effects similar to those of sorbate. The use of 0.05% sodium nitrite with sorbate enhanced the lysis of outgrowing cells at pH 7.2 or lower. Emergence and elongation were inhibited by 0.05% linoleic acid with or without 1.0% sorbate at pH 7.0 to 7.2. The addition of 0.5% tripolyphosphate to media containing 1.5% sorbate at pH 7.1 prevented normal cell growth to an extent greater than with sorbate alone.  相似文献   

15.
A series of catalases have been prepared in which a proportion of the carboxyl groups of glutamate and aspartate residues have been amidated with glycinamide. The physical properties of the amidated catalases have been investigated with specific reference to their interaction with sodium n-dodecyl sulphate (SDS). Amidation leads to an increase in SDS binding at pH 6.4. Microcalorimetric measurements show that the exothermic enthalpy of interaction with SDS increases with the extent of amidation in acid solution (pH 3.2–6.4). The increase in exothermicity is compensated by a decrease in entropy since the average Gibbs energy of SDS binding is independent of the extent of amidation. At pH 3.2 where the catalase carboxyl groups are largely un-ionized amidation still increase the exothermicity of the interaction with SDS. It is suggested that at low pH the SDS anion interacts favourably with the resonance stabilized O-protonated form of amidated side chains.  相似文献   

16.
The factors regulating glucose-6-phosphatase (EC 3.1.3.9) activity and substrate specificity in hepatic microsomes were studied by determining the rate-limiting reaction for the hydrolysis of glucose-6-P, and by examining the effect of detergent activation on phosphotransferase activity. Examination of the pre-steady state kinetics of glucose-6-phosphatase revealed that the steady state rate is determined by the rate of hydrolysis of the enzyme-P intermediate. Treatment of the enzyme with detergent does not alter the extent of the rapid release of glucose per mg of protein, but activates the steady state rate of catalytic turnover. Specificity of the enzyme was evaluated by comparing the effects of mannose and glucose as phosphate acceptors in the phosphotransferase reaction catalyzed by glucose-6-phosphatase. Untreated glucose-6-phosphatase discriminates against mannose as compared with glucose in that mannose and glucose bind to the enzyme-P intermediate of untreated enzyme, but mannose is not an acceptor of Pi. Mannose is an acceptor, however, after treatment of microsomes with detergent. These data cannot be explained in terms of the currently accepted "compartmentation" model for the regulation of glucose-6-phosphatase. The detergent-induced changes in kinetic properties appear to reflect alterations in the intrinsic characteristics of glucose-6-phosphatase, which could result from interaction with its membrane environment.  相似文献   

17.
The techniques of cell electrophoresis and electro-orientation spectroscopy were used to study the effect of sodium dodecyl sulfate (SDS) and cetyl trimethyl ammonium bromide (CTAB) on Escherichia coli K-12 cells from the culture at the exponential and stationary growth phases. SDS (2 x 10(-4) M) considerably damaged cells at the exponential phase, particularly at pH less than 6.0, whereas cells at the stationary phase were damaged to a less degree and only at pH less than 5.3 or after their treatment with Trilon B. The damaging effect of SDS decreased in an isotonic medium (0.25 M sucrose) as compared to a hypotonic medium (distilled water). CTAB also damaged cells at the exponential phase more than those at the stationary phase, and its damaging action decreased with pH. Mg2+, Ca2+, and Sr2+ cations diminished the degree of cell damage with CTAB, but did not exert any noticeable protection in the case of SDS. The different sensitivity of cells at the exponential and stationary growth phases may be associated with changes in their surface electric charge and with the existence of hydrophobic regions on the cell surface. The higher electric charge of cells at the stationary growth phase is presumed to stem from a rise in the amount of surface lipopolysaccharides which bear a negative electric charge.  相似文献   

18.
Guanine deaminase (guanine aminohydrolase, EC 3.5.4.3) from pig brain was purified to homogeneity by column chromatography and ammonium sulphate fractionation. Homogeneity was established by polyacrylamide gel electrophoresis in the presence and absence of sodium dodecyl sulphate (SDS). The molecular weight of 110 000 was determined by gel filtration and sucrose density gradient centrifugation. SDS polyacrylamide gel electrophoresis indicated subunits of a molecular weight of 50 000. The amino acid composition, the isoelectric point and the number of -SH groups were determined. 5.5'-Dithiobis-(2-nitrobenzoic acid) reacts with about seven -SH groups in the native enzyme, but upon denaturation with SDS, 10 -SH groups react with this former reagent. Using electrolytic reduction, 44 half-cystines were determined in accordance with the number of cysteic acid residues determined by amino acid analysis after performic acid oxidation. The Km values determined for substrates of the enzyme were 1.1 . 10(-5) M for guanine in 0.1 M Tris. HCl buffer (pH 8.0) and 3.3 . 10(-4) M for 8-azaguanine in 0.1 M phosphate buffer, pH 6.4. The pKa values determined for ionizable groups of the active site of the enzyme were near pH 6.2 and pH 8.2. The chemical and kinetic evidence suggests that cysteine and histidine may be essential for the catalysis.  相似文献   

19.
Possible reasons are discussed for the abnormal erythrocyte concentration influence on the course of detergent hemolysis curves. The data obtained lead to a conclusion that erythrocyte plasma membrane properties and, as a consequence, the parameters of erythrocyte-detergent interaction are dependent on cell concentration. The dependence is more sharply expressed for concentrations 5.10(7) cell/ml. In the concentration range (5-20) X 10(7) cell/ml the state of erythrocyte membrane remains apparently unchanged.  相似文献   

20.
The effect of surface charge on the porcine pancreatic phospholipase A2 catalyzed hydrolysis of organized substrates was examined through initial rate enzyme kinetic measurements. Two long-chain phospholipid substrates, phosphatidylglycerol (PG) and phosphatidylcholine (PC), were solubilized in seven detergents differing in polar head-group charge. The neutral or zwitterionic detergents selected were Triton X-100, Zwittergent 314, lauryl maltoside, hexadecylphosphocholine (C16PN), and 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate. The negatively and positively charged detergents used were cholate and CTAB, respectively. In general, the negatively charged phospholipid PG was hydrolyzed much more rapidly than the neutral (zwitterionic) phospholipid PC. The rate of hydrolysis of PG was rapid when solubilized in all the neutral detergents and in cholate but was essentially zero in the positively charged CTAB. Conversely, hydrolysis of PC was negligible when solubilized in neutral detergents, except C16PN, and was maximal in the negatively charged detergent, cholate. The rate of hydrolysis of PC solubilized in a neutral detergent became significant only when a negative surface charge was introduced by addition of SDS. Taken together, these kinetic measurements indicate that the surface charge on the lipid aggregates is an important factor in the rate of hydrolysis of phospholipid substrates and the highest activity is observed when the net surface charge is negative. Fluorescence and electron spin resonance (ESR) spectroscopic data provide additional support for this conclusion. The fluorescence emission spectrum of the single tryptophan of phospholipase A2 is a sensitive monitor of interfacial complex formation and shows that interaction of the protein with detergent micelles is strongly dependent on the presence of a negatively charged amphiphile.(ABSTRACT TRUNCATED AT 250 WORDS)  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