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 kinetic parameters of the sugar transport in avian erythrocytes were evaluated under aerobic and anaerobic conditions. In anaerobic cells, transport measurements with 3-O-[14C] methylglucose resulted in a set of similar dissociation-like constants. Thus the Michaelis constants of 3-O-[14C] methylglucose entry and exit, Kso and Ksi, were 8 and 7 mM, respectively. The equilibrium exchange constant, Bs, and the counterflow constant, Rs, were 9 and 11 mM, respectively. The activity constant for 3-O-methylglucose transport, Fs, defined as V/Km, was 4 ml/h per g. This set of kinetic constants was compatible with a symmetrical mobile-carrier model. In contrast, the Michaelis constant for glucose entry, Kgo, was 2 mM and less than the counterflow constant, Rg (8 mM). This result could be accounted for by slower movement of the glucose-carrier complex than the free carrier. The activity constant for glucose transport, Fg, was 5 ml/h perg.Under aerobic conditions, two of the dissociation-like constants (Ksi and Bs) for 3-O-methylglucose transport were significantly larger than those obtained in anaerobic cells, but the remaining two (Kso and Rs) remained unchanged. The values, for Kso, Ksi, Bs and Rs were 8, 26, 20 and 8 mM, respectively. The activity constant, Fs, decreased to 2 ml/h per g. These changes in kinetic constants were consistent with the hypothesis that anoxia accelerated sugar transport by releasing free carrier that was previously sequestered on the inside of the cell membrane.  相似文献   

2.
The addition of cholate to the microsomes at 37.5°C resulted in a striking decrease in the apparent substrate dissociation constant (K′s) and its temperature dependency. The microsomal membranes depleted of 80% of the lipids preserved the temperature dependency of the Ks and exhibited breaks in the Van't Hoff plot at the characteristic temperature of the lipids phase transition. The results indicate that the cytochrome P-450 is considerably restrained from expressing its maximum substrate binding potential at physiological temperature. In addition, the results indicate that the majority of the lipids apparently do not play a significant role in imposing constraint on the substratecytochrome P-450 binding reaction and in the temperature dependency of the Ks.  相似文献   

3.
A number of algebraic expressions for the solute concentration for use with the Michaelis-Menten equation during the analysis of data from intestinal perfusion experiments have been investigated. It is concluded that the most suitable, especially if water absorption is occuring, is of the form: s=(sinitial ? sefluent)/In(sinitial/seffluent).  相似文献   

4.
(1) Treatment of (Na+ + K+)-ATPase from rabbit kidney outer medulla with the γ-35S labeled thio-analogue of ATP in the presence of Na+ + Mg2+ and the absence of K+ leads to thiophosphorylation of the enzyme. The Km value for [γ-S]ATP is 2.2 μM and for Na+ 4.2 mM at 22°C. Thiophosphorylation is a sigmoidal function of the Na+ concentration, yielding a Hill coefficient nH = 2.6. (2) The thio-analogue (Km = 35 μM) can also support overall (Na+ + K+)-ATPase activity, but Vmax at 37°C is only 1.3 γmol · (mg protein)? · h?1 or 0.09% of the specific activity for ATP (Km = 0.43 mM). (3) The thiophosphoenzyme intermediate, like the natural phosphoenzyme, is sensitive to hydroxylamine, indicating that it also is an acylphosphate. However, the thiophosphoenzyme, unlike the phosphoenzyme, is acid labile at temperatures as low as 0°C. The acid-denatured thiophosphoenzyme has optimal stability at pH 5–6. (4) The thiophosphorylation capacity of the enzyme is equal to its phosphorylation capacity, indicating the same number of sites. Phosphorylation by ATP excludes thiophosphorylation, suggesting that the two substrates compete for the same phosphorylation site. (5) The (apparent) rate constants of thiophosphorylation (0.4 s?1 vs. 180 s?1), spontaneous dethiophosphorylation (0.04 s?1 vs. 0.5 s?1) and K+-stimulated dethiophosphorylation (0.54 s?1 vs. 230 s?1) are much lower than those for the corresponding reactions based on ATP. (6) In contrast to the phosphoenzyme, the thiophosphoenzyme is ADP-sensitive (with an apparent rate constant in ADP-induced dethiophosphorylation of 0.35 s?1, KmADP = 48 μM at 0.1 mM ATP) and is relatively K+-insensitve. The Km for K+ in dethiophosphorylation is 0.9 mM and in dephosphorylation 0.09 mM. The thiophosphoenzyme appears to be for 75–90% in the ADP-sensitive E1-conformation.  相似文献   

5.
Binding of the structural protein soc to the head shell of bacteriophage T4   总被引:5,自引:0,他引:5  
Qβ plus strands with a 70 S ribosome bound to the coat cistron initiation site were used as template for Qβ replicase. Minus strand synthesis proceeded until the replicase reached the ribosome. The ribosome was removed and elongation was continued in a substrate-controlled, stepwise fashion. The nucleotide analog N4-hydroxyCMP was introduced into the positions complementary to the third and fourth nucleotides of the coat cistron. The minus strands were elongated to completion, purified and used as template for Qβ replicase. The final plus strand preparation consisted of four species, with the sequences -A-U-G-G- (wild type), -A-U-A-G- (mutant C3), -A-U-G-A- (mutant C4) and -A-U-A-A- (mutant C3C4) at the coat initiation site. The ribosome binding capacity of the mutant RNAs relative to wild type was <0.1 (C3), 3.2 (C4) and 0.3 (C3C4). The finding that mutant C3 no longer formed an initiation complex suggests that the interaction of the ribosome binding site with fMet-tRNA plays an essential role in the formation of the 70 S initiation complex. The fact that mutant C4 RNA bound more efficiently than wild type, and that mutant C3C4 RNA showed substantial ribosome binding capacity whereas the single mutant C3 did not, can be explained by assuming that an A residue following the A-U-G triplet interacts with a complementary U residue in the anticodon loop sequence. In the case of C3C4 this additional base-pair may offset the reduced codon-anticodon interaction resulting from the modification of the A-U-G codon.  相似文献   

6.
In an accompanying publication by Duckwitz-Peterlein, Eilenberger and Overath ((1977) Biochim. Biophys. Acta 469, 311–325) it is shown that the exchange of lipid molecules between negatively charged vesicles consisting of total phospholipid extracts from Escherichia coli occurs by the transfer of single lipid monomers or small micelles through the water. Here a kinetic interpretation is presented in terms of a rate constant, k?, for the escape of lipid molecules from the vesicle bilayer into the water. The evaluated rate constants are k?P = (0.86 ± 0.05) · 10?5s?1 and k?E = (1.09 ± 0.13) · 10?6s?1 for phospholipid molecules with trans-Δ9-hexadecenoate and trans-Δ9-octadecenoate, respectively, as the predominant acyl chain component. The rate constants are discussed in terms of the acyl chain and polar head group composition of the lipids.  相似文献   

7.
The uptake of d-glucose, 2-aminoisobutyric acid and glycine was studied with intestinal brush border membrane vesicles of a marine herbivorous fish: Boops salpa. The uptake of these three substances is stimulated by an Na+ electrochemical gradient (CoutCin). For glucose, an increase of the electrical membrane potential generated by a concentration gradient of the liposoluble anion, SCN?, increases the Na+-dependent transport. This responsiveness to the membrane potential was confirmed by valinomycin. Differently from glucose, uptake of glycine and 2-aminoisobutyric acid requires, besides the Na+ gradient, the presence of Cl? on the external side of the vesicles. In the absence of Cl?, amino acid uptake is not stimulated by the Na+ gradient and is not influenced by an electrical membrane potential generated by SCN? gradient (Cout>Cin) or by a K+ diffusion potential (Cin>Cout). This Cl? requirement differs from the Na+ requirement, since a Cl? gradient (Cout>Cin) does not result in an accumulation of glycine or 2-aminoisobutyric acid similar to that produced by an Na+ gradient.  相似文献   

8.
Pyrene lecithin, a new excimer-forming lipid molecule, has been synthesized to examine the transversal mobility of probe molecules in lecithin bilayer vesicles. The rate of the lipid exchange is obtained by following the excimer yield as a function of time after mixing of fluorescence doped and undoped vesicles. A rapid exchange (τ12 = 11 s) is followed by a slow transfer (t12 = 8 h). Above the lipid phase transition the fast transfer can be attributed to an exchange of lipid molecules from the outer layer of one vesicle to the outer layer of another one. The slow exchange is interpreted in terms of the ‘flip-flop’ process between the two layers of a single bilayer vesicle.Using pyrene and pyrene decanoic acid as probe molecules only the fast transfer through the water phase is observed (τ12 = 4 s for pyrene and τ12 = 7 s for pyrene decanoic acid). This indicates that molecules like fatty acids or apolar membrane constituents must equilibrate very rapidly in a single bilayer vesicle.The water solubility or the critical micelle concentration of the probe molecules is determined and related to the transfer rates. An exchange process through the water phase via a monomeric state can be excluded.  相似文献   

9.
10.
11.
If the bicyclic peptide ring proposed by Gross etal. (1,2) does in fact exist in nisin and related antibiotics, then the unusual β-methyllanthionine component must be significantly distorted from its conformation in the free state, as determined by x-ray structure analysis. The torsion angles about the SCβ bonds are 50–100° from the torsion angles in models of the sulfur-bridged peptide ring proposed for nisin. The chirality of the methylated β-carbon atom is (S). The conformation of the amino acid differs from that of meso-lanthionine only by a 180° rotation of a carboxyl group about the CαDCβ(CH3) bond.  相似文献   

12.
《Inorganica chimica acta》1986,115(2):169-172
2-(Methylamino)pyridine reacts with RuCl2(CO)3 to give a carbamoyl complex, [Ru(C(O)N(CH3)(C5H4N)Cl(CO)2], which yields with pyridine (py) and acetylacetone (Hacac), respectively, [Ru(C(O)N(CH3)C5H4N)Cl(CO)2(py)] and [Ru(C(O)N(CH3)C5H4N)(CO)2(acac)]. These complexes are characterized spectroscopically. The amino group of the ligand is carbonylated and the resulted carbamoyl ligand is chelating through a pyridine ring-N and a carbamoyl-C atom. 2-Aminopyridine and 2-aminopyrimidine react similarly with RuCl2(CO)3 to give the corresponding carbamoyl complexes.  相似文献   

13.
Single spherical bilayer membranes of the Pagano-Thompson type (Pagano, R. and Thompson, T.E. (1967) Biochim. Biophys. Acta 144, 666–669), formed from monooleyl phosphate and cholesterol dissolved in CHCl3/CH3OH/n-decane, were subjected to a fast impedance analysis of high precision. Dielectric behavior of the whole system, as monitored from outside the spherical membrane, was sensitive to changes in the membrane state from the thick colored to the thin black state. With a spherical membrane 2–3 mm in diameter formed in the sample cavity containing 0.12 ml 10 mM NaCl, the former state was characterized by a dielectric dispersion having dielectric increment (Δ?) of some 102 and characteristic frequency (?c) around 106 Hz, while the latter had Δ? ? 105and ?c ? 103Hz. Complex plane plots for both dispersions traced semicircles, indicating that the present system may be unequivocally analyzed to yield spherical radius and membrane capacity (Cm) on the basis of a well-established dielectric theory. Cm for the thin membranes has thus been determined to be 0.54 μF · cm?2, in excellent agreement with a separate determination on planar membranes. The applicability of the present type of spherical membranes under dielectric monitoring to the study of membrane fusion or of exocytosis is suggested.  相似文献   

14.
15.
(1) The polymorphic phase behaviour of aqueous dispersions of various synthetic phosphatidylethanolamines, both singly and in mixtures, has been investigated by 31P-NMR. (2) 14:014:0 PE remains in the lamellar phase up to 90°C. 18:1t18:1t PE exhibits a lamellar to hexagonal (HII) transition between 60°C and 63°C. For 18:1c18:1c PE, the lamellar to hexagonal (HII) transition occurs between 7 and 12°C, whereas for 18:2c18:2c PE, the hexagonal (HII) phase is the preferred structure above ?15°C. (3) Mixtures of 18:1c18:1c PE and 18:1t18:1t PE exhibit near-ideal miscibility behaviour. For mixtures of 18:1c18:1c PE and 14:014:0 PE there is evidence of fluid-solid immiscibility at temperatures below the gel-liquid crystalline transition temperature of the 14:014:0 PE component. Mixtures of 18:2c18:2c PE and 18:1t18:1t PE exhibit complex phase behaviour involving limited fluid-solid immiscibility at low temperatures and formation of a phase allowing isotropic motional averaging at higher temperatures. (4) 31P-NMR provides a graphic method for investigating the miscibility properties of mixed PE systems.  相似文献   

16.
In 1976 (Horton, A.W., Butts, C.K. and Schuff, A.R. (1976) Colloid Interface Sci. 5, 159–168) we assayed pristance (2,6,10,14-tetramethylpentadecane) in a model interfacial system that has given excellent correlation with cocarcinogenic activity among n-alkanes, as tested in cycloalkane diluents. It was predicted that this branched-chain derivative of the diterpenes would have higher activity than n-C18H38, one of the most cocarcinogenic of the n-alkanes in such diluents. Pristane was compared with n-C18H38 and two other n-alkanes for their promoting activities in cyclohexane for C3H male mice after a single application of 7,12-dimethylbenz[a]anthracene. The branched-chain alkane proved to be more active. 20% n-tetracosane in cyclohexane was inactive, which correlated with its effects in this diluent in the physical assay system. The promoting activity of 75% n-octane in cyclohexane, predicted by the physical assay, was confirmed by tests on mice. The combined by-products of the synthesis of tetracosane, including C12 alkanes and alkenes, C19 and C20 alkylbenzenes, and C24 alkenes, proved to be a very active promoter. However, a mixture in cyclohexane of purified tetracosane with this composite of potential impurities was inactive. From the alkanes behavior in physical systems involving vatious membrane phospholipids and steroids, it is hypothesized that the primary step in their biological activity is a chain-chain interaction with membrane lipids that alters the properties of liquid-crystalline phases at aqueous interfaces. Resulting changes in the microfluidity of the lipid phase and the lateral mobility of critical hormone receptors and enzyme systems, such as the nucleotidyl cyclases, would perturb control systems that maintain the normal behavior of the initiated cell. Thus, its progression to a proliferating neoplasm may be favored.  相似文献   

17.
The natural affinity of various bacterial glycopeptides and lipopolysaccharides for mammalian cell membranes was estimated quantitatively by comparison with the adsorption of lipopolysaccharide from Escherichia coli NCTC 8623 to erythrocytes, thymocytes, bone marrow cells, spleen cells, peritoneal lymphocytes and macrophages. Immunopotentiating activity was estimated by measuring the ability of the bacterial fractions to stimulate a humoral response to ovalbumin in HAM/1CR mice. When the affinity for mammalian cell membranes was compared with the stimulation of the antibody response, it was found that a negative correlation for peritoneal macrophages (rs=?0.94, P<0.0005) and a positive correlation for peritoneal lymphocytes (rs=+0.97,P<0.0005) and spleen cells (rs=+0.76,P<0.005) existed.  相似文献   

18.
Catalytic activity of thymidylate synthase, as measured in, vivo, is tightly linked to S phase of the cell cycle in Chinese hamster embryo fibroblast cells. This activity, as measured in, vitro, is found in all parts of the cell cycle. Thymidylate synthase activity in nuclear (karyoplast) extracts increased as the cells progressed from G0G1 to S phase. This enzymatic activity in the nuclei of S phase cells is associated with the multienzyme complex (replitase) that also contained DNA polymerase and other enzymes of DNA replication and precursor synthesis. The degree of association of thymidylate synthase with replitase, which increased co-ordinately as the cells progressed from G0G1 phase to S phase, coincided strongly with the level of in, vivo activity of the enzyme.  相似文献   

19.
Soluble (Na++K+)-ATPase consisting predominantly of αβ-units with Mr below 170 000 was prepared by incubating pure membrane-bound (Na++K+)-ATPase (35–48 μmol Pi/min per mg protein) from the outer renal medulla with the non-ionic detergent dodecyloctaethyleneglycol monoether (C12E8). (Na++K+)-ATPase and potassium phosphatase remained fully active in the detergent solution at C12E8/protein ratios of 2.5–3, at which 50–70% of the membrane protein was solubilized. The soluble protomeric (Na++K+)-ATPase was reconstituted to Na+, K+ pumps in phospholipid vesicles by the freeze-thaw sonication procedure. Protein solubilization was complete at C12E8/protein ratios of 5–6, at the expense of partial inactivation, but (Na++K+)-ATPase and potassium phosphatase could be reactivated after binding of C12E8 to Bio-Beads SM2. At C12E8/protein ratios higher than 6 the activities were irreversibly lost. Inactivation could be explained by delipidation. It was not due to subunit dissociation since only small changes in sedimentation velocities were seen when the C12E8/protein ratio was increased from 2.9 to 46. As determined immediately after solubilization, S20,w was 7.4 S for the fully active (Na++K+)-ATPase, 7.3 S for the partially active particle, and 6.5 S for the inactive particle at high C12E8/protein ratios. The maximum molecular masses determined by analytical ultracentrifugation were 141 000–170 000 dalton for these protein particles. Secondary aggregation occurred during column chromatography, with formation of enzymatically active (αβ)2-dimers or (αβ)3-trimers with S20,w=10–12 S and apparent molecular masses in the range 273 000–386 000 daltons. This may reflect non-specific time-dependent aggregation of the detergent micelles.  相似文献   

20.
The rates of electron exchange between ferricytochrome c (CIII)3 and ferrocytochrome c (CII) were observed as a function of the concentrations of ferrihexacyanide (FeIII) and ferrohexacyanide (FeII) by monitoring the line widths of several proton resonances of the protein. Addition of FeII to CIII homogeneously increased the line widths of the two downfield paramagnetically shifted heme methyl proton resonances to a maximal value. This was interpreted as indicating the formation of a stoichiometric complex, CIII·FeII, in the over-all reaction:
CIII+FeII?k?1k1CIII·FeII?k?2k2CII·FeIII?k?3k3CIII+FeII
Values for k1k?1 = 0.4 × 103m?1and k2 = 208 s?1, respectively, were calculated from the maximal change in line width observed at pH 7.0 and 25 °C. Changes in the line width of CIII in the presence of FeII and either KCl or FeIII suggest that complexation is principally ionic, that FeIII and FeII compete for a common site. Addition of saturating concentrations of FeIII to CIII produced only minor changes in the nuclear magnetic resonance spectrum of CIII suggesting that complexation occurs on the protein surface.Addition of FeIII to CII in the presence of excess FeII (to retain most of the protein as CII) increased the line width of the methyl protons of ligated methionine 80. A value for k?2 ≈ 2.08 × 104 s?1 was calculated from the dependence of linewidth on the concentration of FeII at 24 °C. These rates are shown to be consistent with the over-all rates of reduction and oxidation previously determined by stopped flow measurements, indicating that k2 and k?2 were rate limiting. From the temperature dependence the enthalpies of activation are 7.9 and 15.2 kcal/mol for k2 and k?2, respectively.  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