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 penicillin-binding protein that is thought to be the lethal target of penicillin in Bacillus megaterium (protein 1) has been purified to greater than 95% homogeneity. The membrane-bound penicillin-binding proteins were solubilized with a non-ionic detergent and partially separated from each other by ion-exchange chromatography on DEAE-Sepharose CL-6B. Protein 1 was subsequently purified by covalent affinity chromatography on ampicillin-affinose. Bacillus licheniformis contains an equivalent penicillin-binding protein (protein 1) that can be more readily purified to virtual homogeneity in a one-step procedure. It was separated from the other penicillin-binding proteins by utilizing the observation that in this organism, this particular protein is the only one whose covalent complex with benzylpenicillin subsequently breaks down. Membranes were treated with saturating concentrations of benzylpenicillin followed by the removal of free penicillin and further incubation to allow the complex between benzylpenicillin and protein 1 to break down. The penicillin-binding proteins were then solubilized and applied to a column of ampicillin-affinose to which only protein 1 was bound as the other penicillin-binding proteins still had benzylpenicillin bound to them. Pure protein 1 was eluted from the affinity resin with hydroxylamine. The interaction of benzylpenicillin with purified protein 1 has been studied by separating unbound antibiotic from the benzylpenicillin . protein complex by paper electrophoresis. Benzylpenicillin reacts with the protein rapidly to form a covalent complex and the fully saturated complex has a molar ratio of bound [14C] benzylpenicillin: protein of 0.7:1. The complex breaks down, obeying first-order kinetics, with a half-life of 16 min at 35 degrees C, a value identical to that obtained with the membrane-bound protein. The concentration of benzylpenicillin that results in the formation of 50% of the maximum amount of benzylpenicillin . protein complex is that at which the molar amount of benzylpenicillin present is equal to 50% of the molar amount of penicillin-binding protein, rather than being a measure of any of the kinetic parameters of the binding reaction. This observation may be significant in the interpretation of previous results where the amounts of penicillins needed to kill cells or to inhibit penicillin-sensitive reactions have been expressed as concentrations. The possible importance of the breakdown of beta-lactam . protein complexes in the clinical use of these antibiotics is discussed.  相似文献   

2.
Thermal destruction processes in proteins involving cystine residues   总被引:5,自引:0,他引:5  
To ascertain the upper limit of protein thermostability and to evaluate the introduction of additional -S-S- bridges for the enhancement of protein thermostability by site-directed mutagenesis, the stability of cystine residues at 100 degrees C has been investigated in a number of unrelated proteins. All proteins studied (more than a dozen) undergo heat-induced beta-elimination of cystine residues in the pH range from 4 to 8. The time courses of this process closely follow first-order kinetics indicating that the stability of a cystine residue is not significantly affected by its position in the polypeptide chain. Furthermore, the first-order rate constants of beta-elimination of disulfide bonds at 100 degrees C are remarkably similar for different proteins--0.8 +/- 0.3 h-1 and 0.06 +/- 0.02 h-1 at pH 8.0 and 6.0, respectively. Thus, this process is relatively independent of both the primary structure and the elements of higher structures remaining in proteins in 100 degrees C. beta-Elimination of disulfides produces free thiols that cause yet another deleterious reaction in proteins--heat-induced disulfide interchange. This reshuffling of -S-S-bonds, which is much faster than beta-elimination, has also been quantitatively characterized.  相似文献   

3.
This study was designed to establish the properties of liver plasma membranes (LPM) Na+,K+-ATPase in the hamster and to determine whether a similar assay may be used to measure enzyme activity in the hamster and in the rat. Maximal Na+,K+-ATPase activity was obtained when the assay medium contained 5 mM Mg APT2- with or without 1 mM free Mg2+, 120 mM Na+, 12,5 mM K+. The incubation must be performed at 37 degrees C, pH 7.4. In the absence of free Mg2+, the saturation curve with respect to the substrate Mg ATP2- resulted in biphasic complex kinetics with a maximal activity at a substrate concentration of 5 mM. In the presence of 1 mM free Mg2+ activation of Na+,K+-ATPase and modification of the kinetics were observed: the biphasic curve tended to disappear and to become of the Michaelis-Menten type. The apparent Km for Mg APT2- was 0.36 mM and the Vmax 34.5 mumol.h-1.mg protein-1. In the presence of 10 mM free Mg2+ a decrease in the Vmax was observed without any effect on the apparent Km for Mg APT2-. It is concluded that the same incubation medium may be used to assay LPM N+,K+-ATPase from hamster and rat and that the addition of 1 mM free Mg2+ to the incubation medium is recommended to obtain Michaelis-Menten kinetics in order to eliminate complex kinetics due to the absence of free Mg2+.  相似文献   

4.
The kinetics of oxidation of iota- and lambda-carrageenan as sulfated carbohydrates by permanganate ion in aqueous perchlorate solutions at a constant ionic strength of 2.0 mol dm−3 have been investigated spectrophotometrically. The pseudo-first-order plots were found to be of inverted S-shape throughout the entire courses of reactions. The initial rates were found to be relatively slow in the early stages, followed by an increase in the oxidation rates over longer time periods. The experimental observations showed first-order dependences in permanganate and fractional first-order kinetics with respect to both carrageenans concentration for both the induction and autoacceleration periods. The results obtained at various hydrogen ion concentrations showed that the oxidation processes in these redox systems are acid-catalyzed throughout the two stages of oxidation reactions. The added salts lead to the prediction that MnIII is the reactive species throughout the autoacceleration periods. Kinetic evidence for the formation of 1:1 intermediate complexes was revealed. The kinetic parameters have been evaluated and tentative reaction mechanisms in good agreement with the kinetic results are discussed.  相似文献   

5.
The presence of a thiol in the steroid binding cavity of glucocorticoid receptors has recently been proved by our affinity labeling of Cys-656 in the steroid binding domain of rat receptors (Simons, S. S., Jr., Pumphrey, J. G., Rudikoff, S., and Eisen, H. J. (1987) J. Biol. Chem. 262, 9676-9680). Studies with the sterically small, thiol-specific reagent methyl methanethiolsulfonate (MMTS) now reveal the involvement of at least two sulfhydryl groups in steroid binding. While the dose-response curves for [3H]dexamethasone binding versus thiol reagent are normally sigmoidal, an unusual bimodal curve is obtained with MMTS in which dexamethasone binding is eliminated at low, but maintained at intermediate, MMTS concentrations. This bimodal dose-response curve demands the involvement of two (or more) thiol groups. Those receptors pretreated with intermediate concentrations of MMTS retain approximately 70% of the initial binding capacity and one-fifth the affinity for dexamethasone. Solutions of this low affinity form of receptor contain essentially no accessible -SH groups, and all of the usual covalent labeling by dexamethasone 21-mesylate of various proteins, including the receptor, is blocked. The facts, that this low affinity form of the receptor is not affected by added iodoacetamide, cannot be produced from the nonsteroid binding form of receptor simply by adding more MMTS, and displays different kinetics of formation than does the nonsteroid binding form of receptor all argue that reaction of the receptor with intermediate and low MMTS, concentrations occurs via different pathways. Nevertheless, the effects of both concentrations of MMTS on the receptor are fully reversible with added dithiothreitol. The kinetics of inhibition of [3H]dexamethasone binding at low MMTS concentrations are independent of receptor concentration, indicating an intramolecular reaction. Collectively these data suggest a model of steroid binding involving two thiols, one of which appears to be Cys-656. Low concentrations of MMTS induce the formation of an intramolecular disulfide, which prevents steroid binding, while the intermediate MMTS concentrations convert both thiols directly to mixed disulfides, and steroid binding persists. Thus, reduced thiols do not appear to be required for steroid binding if the steric bulk of the oxidized thiols is small.  相似文献   

6.
Partially purified ornithine decarboxylase, isolated from the liver of thioacetamide-treated rats, is stable in the absence of added low-molecular-mass thiols or other reducing agents. However, under these conditions, the enzyme is rapidly inactivated upon incubation with L-ornithine or L-2-methylornithine. The inactivation process follows first-order kinetics, and saturation kinetics are observed. Rapid recovery of activity is observed after subsequent addition of dithiothreitol. As distinct from L-ornithine, D-ornithine, putrescine, spermidine, or spermine do not produce inactivation of ornithine decarboxylase. Very similar results are obtained with pure ornithine decarboxylase isolated from androgen-stimulated mouse kidney, stabilized with a rat liver extract.  相似文献   

7.
The pairing of single- and double-stranded DNA molecules at homologous sequences promoted by recA and single-stranded DNA-binding proteins of Escherichia coli follows apparent first-order kinetics. The initial rate and first-order rate constant for the reaction are maximal at approximately 1 recA protein/3 and 1 single-stranded DNA-binding protein/8 nucleotides of single-stranded DNA. The initial rate increases with the concentration of duplex DNA; however, the rate constant is independent of duplex DNA concentration. Both the rate constant and extent of reaction increase linearly with increasing length of duplex DNA over the range 366 to 8623 base pairs. In contrast, the rate constant is independent of the size of the circular single-stranded DNA between 6,400 and 10,100 nucleotides. No significant effect on reaction rate is observed when a single-stranded DNA is paired with 477 base pairs of homologous duplex DNA joined to increasing lengths of heterologous DNA (627-2,367 base pairs). Similarly, heterologous T7 DNA has no effect on the rate of pairing. These findings support a mechanism in which a recA protein-single-stranded DNA complex interacts with the duplex DNA to produce an intermediate in which the two DNA molecules are aligned at homologous sequences. Conversion of the intermediate to a paranemic joint then occurs in a rate-determining unimolecular process.  相似文献   

8.
The time-dependent absorbance change that occurs when benzo[alpha]pyrene 7,8-diol-9,10-epoxide is added to solutions of calf thymus DNA has been shown, by an unequivocal chromatographic method, to correspond to DNA-catalyzed hydrolysis of the diol-epoxide. At 25 degrees C and mu = 0.10, the kinetics of the reaction of the diol-epoxide with polyadenylic acid or DNA are consistent with preequilibrium formation of a non-covalent complex between the diol-epoxide and the polynucleotide or DNA, followed by hydrolysis of the bound epoxide by a process that is first-order in hydronium ions. Cacodylic acid also catalyzes the hydrolysis of the epoxide bound to polyadenylic acid. The rate of the DNA-catalyzed hydrolysis exhibits little or no enantiomeric selectivity for the diol-epoxide. DNA catalyzed hydrolysis of the diol-epoxide is extraordinarily sensitive to the salt concentration in the reaction medium: the rate of hydrolysis of the bound epoxide at pH 7 is retarded by a factor of approximately 45 in the presence of 0.1 M sodium chloride compared to a 1 mM buffer containing no added salt. Thus, studies of the interactions of DNA with carcinogenic diol-epoxides must take into account the ionic environment of DNA within the cell.  相似文献   

9.
Members of the human heat shock (HSP) family of related proteins are involved in the intracellular folding, transport, and assembly of proteins and protein complexes. We have observed that human heat shock protein 70 (HSP70) is associated with the capsid precursor P1 of poliovirus and coxsackievirus B1 in infected HeLa cells. Antiserum generated against HSP70 coimmunoprecipitated the poliovirus protein P1, an intermediate in capsid assembly. Similarly, alpha-virion serum coimmunoprecipitated HSP70 from virus-infected cell extracts, but not from mock-infected cell extracts. The HSP70-P1 complex was stable in high-salt medium but was sensitive to incubation with 2 mM ATP, which is a characteristic of other known functional complexes between HSP70 and cellular proteins. The P1 in the complex was predominantly newly synthesized, and the half-life of complexed P1 was nearly twice as long as that of total P1. The HSP70-P1 complex was found to sediment at 3S to 6S, suggesting that it may be part of, or a precursor to, the "5S promoter particles" thought to be an assembly intermediate of picornaviruses. The finding that HSP70 was associated with the capsid precursors of at least two enteroviruses may suggest a functional role of these complexes in the viral life cycles.  相似文献   

10.
Fructose 1,6-bisphosphate aldolase from rabbit muscle forms by reaction with dihydroxyacetone phosphate a pyruvaldehyde-aldolase-orthophosphate complex that is in equilibrium with the eneamine intermediate. The new intermediate accumulates in two phases. The first one is practically complete in 40ms, and the second occurs with an apparent first-order rate constant of 4.6 +/- 0.5s-1. The new intermediate breaks down slowly with the release into the medium of pyruvaldehyde and Pi. The rate of the spontaneous release is higher at acidic than at neutral pH.  相似文献   

11.
Dimethylglycine dehydrogenase (EC 1.5.99.2) and sarcosine dehydrogenase (EC 1.5.99.1) are flavoproteins which catalyze the oxidative demethylation of dimethylglycine to sarcosine and sarcosine to glycine, respectively. During these reactions tightly bound tetrahydropteroylpentaglutamate (H4PteGlu5) is converted to 5,10-methylene tetrahydropteroylpentaglutamate (5,10-CH2-H4PteGlu5), although in the absence of H4PteGlu5, formaldehyde is produced. Single turnover studies using substrate levels of the enzyme (2.3 microM) showed pseudo-first-order kinetics, with apparent first-order rate constants of 0.084 and 0.14 s-1 at 23 and 48.3 microM dimethylglycine, respectively, for dimethylglycine dehydrogenase and 0.065 s-1 at 47.3 microM sarcosine for sarcosine dehydrogenase. The rates were identical in the absence or presence of bound tetrahydropteroylglutamate (H4PteGlu). Titration of the enzymes with substrate under anaerobic conditions did not disclose the presence of an intermediate semiquinone. The effect of dimethylglycine concentration upon the rate of the dimethylglycine dehydrogenase reaction under aerobic conditions showed nonsaturable kinetics suggesting a second low-affinity site for the substrate which increases the enzymatic rate. The Km for the high-affinity active site was 0.05 mM while direct binding for the low-affinity site could not be measured. Sarcosine and dimethylthetin are poor substrates for dimethylglycine dehydrogenase and methoxyacetic acid is a competitive inhibitor at low substrate concentrations. At high dimethylglycine concentrations, increasing the concentration of methoxyacetic acid produces an initial activation and then inhibition of dimethylglycine dehydrogenase activity. When these compounds were added in varying concentrations to the enzyme in the presence of dimethylglycine, their effects upon the rate of the reaction were consistent with the presence of a second low-affinity binding site on the enzyme which enhances the reaction rate. When sarcosine is used as the substrate for sarcosine dehydrogenase the kinetics are Michaelis-Menten with a Km of 0.5 mM for sarcosine. Also, methoxyacetic acid is a competitive inhibitor of sarcosine dehydrogenase with a Ki of 0.26 mM. In the absence of folate, substrate and product determinations indicated that 1 mol of formaldehyde and of sarcosine or glycine were produced for each mole of dimethylglycine or sarcosine consumed with the concomitant reduction of 1 mol of bound FAD.  相似文献   

12.
The kinetics of esterification of exogenous retinol by cell membranes prepared from the crude homogenate of the frog retinal pigment epithelium was studied. The formation of retinyl palmitate from added retinol was directly assayed by high performance liquid chromatography (HPLC). A linear relationship was observed between the amount of protein (up to 2 mg) in the incubation medium and the amount of retinyl palmitate formed. At room temperature, this reaction took less than 2 hours to complete. By varying the substrate concentration in the incubation medium, the reciprocal of initial velocity of the reaction (nmol retinyl palmitate formed per hour) was plotted against the reciprocal of substrate concentration (nmol of retinol). This double-reciprocal plot shows that the apparent Km of the reaction was 10 microM with an apparent Vmax of 9.1 nmol of retinyl palmitate per hour per mg protein. When this assay was repeated in the presence of 3,4-didehydroretinol (20 microM), the kinetics of the reaction showed the pattern of that of a competitive inhibitor, suggesting that 3,4-didehydroretinol competes with retinol for the same active site for esterification. The esterification of 3,4-didehydroretinol resulted in the formation of 3,4-didehydroretinyl palmitate, which was also measured by HPLC. The amount of 3,4-didehydroretinyl palmitate formed by this reaction decreased in proportion to increased retinol concentration in the incubation mixture. This further confirms that a competition exists between the esterification of retinol and 3,4-didehydroretinol by retinal pigment epithelium of the frog.  相似文献   

13.
The ultrastructural localization of glucose-6-phosphate dehydrogenase (NADP-linked) has been attempted in steroid-secreting cells. Rat adrenocortical cells and newt testicular glandular cells were fixed in an ice-cold mixture of 1% methanol-free formaldehyde and 0.25% glutaraldehyde. Potassium ferricyanide was used as the final electron acceptor. After incubation, the final copper ferrocyanide precipitate is exclusively observed in the hyaloplasm of these cells, provided that an electron carrier (1.0 mM PMS) has been added to the medium in order to by-pass the tissue "diaphorase" (NADPH-ferricyanide reductase) reaction. No precipitate appears in the absence of glucose-6-phosphate (substrate). Incubation in a medium devoid of PMS results in an exclusively mitochondrial reaction; the latter is that of the "diaphorase", which in these cells is mitochondrial. These results prove the importance of utilizing exogenous electron carriers (such as PMS) in coenzyme-linked dehydrogenase cytochemistry. Although polyvinyl alcohol was included in the washing and incubation media, in order to increase their viscosity, problems still exist concerning ultracytochemical localization of this "soluble" enzyme; these problems are discussed in the paper.  相似文献   

14.
The kinetics of photolysis of riboflavin (RF) in water (pH 7.0) and in organic solvents (acetonitrile, methanol, ethanol, 1-propanol, 1-butanol, ethyl acetate) has been studied using a multicomponent spectrometric method for the assay of RF and its major photoproducts, formylmethylflavin and lumichrome. The apparent first-order rate constants (kobs) for the reaction range from 3.19 (ethyl acetate) to 4.61 × 10−3 min−1 (water). The values of kobs have been found to be a linear function of solvent dielectric constant implying the participation of a dipolar intermediate along the reaction pathway. The degradation of this intermediate is promoted by the polarity of the medium. This indicates a greater stabilization of the excited-triplet states of RF with an increase in solvent polarity to facilitate its reduction. The rate constants for the reaction show a linear relation with the solvent acceptor number indicating the degree of solute–solvent interaction in different solvents. It would depend on the electron-donating capacity of RF molecule in organic solvents. The values of kobs are inversely proportional to the viscosity of the medium as a result of diffusion-controlled processes.KEY WORDS: dielectric constant, kinetics, photolysis, riboflavin, solvent effect, viscosity  相似文献   

15.
In the presence of ATP hepatic microsomes sequester calcium. This sequestration is thought to be important in the modulation of free cytosolic calcium concentration. We find that on the addition of NADPH the uptake of calcium by the hepatic microsomes is inhibited 27-85%. This inhibition is reversed by the addition of 1 mM reduced glutathione (85-91% of control), incubation under a nitrogen atmosphere (112% of control), or incubation in a 80% carbon monoxide/20% oxygen atmosphere (75% of control). Superoxide dismutase had no effect on the inhibition, while catalase reversed the inhibition by 35%. The addition of 1 mM reduced glutathione at 2 and 5 min after the addition of NADPH led to uptakes of calcium which paralleled the uptake seen when the reduced glutathione was added at the beginning of the incubation. The effect of reduced glutathione showed saturation kinetics with a Km of 10 microM. Together these data suggest that cytochrome P-450 reduces the activity of the microsomal ATP-dependent calcium pump both by the production of hydrogen peroxide and by the direct oxidation of the protein thiols. The reversal of this effect by reduced glutathione appears to be enzymatically catalyzed.  相似文献   

16.
1. The kinetics of ferrocytochrome c peroxidation by yeast peroxidase are described. Kinetic differences between the older and more recent preparations of the enzyme most probably arise from differences in intrinsic turnover rates. 2. The time-courses of cytochrome c peroxidation by the enzyme follow essentially first-order kinetics in phosphate buffer. Deviations from first-order kinetics occur in acetate buffer, and are due to a higher enzymic turnover rate in this medium accompanied by a greater tendency to autocatalytic peroxidation of cytochrome c. 3. The kinetics of ferrocytochrome c peroxidation by yeast peroxidase are interpreted in terms of a mechanism postulating formation of reversible complexes between the peroxidase and both reduced and oxidized cytochrome c. Formation of these complexes is inhibited at high ionic strengths and by polycations. 4. Oxidized cytochrome c can act as a competitive inhibitor of ferrocytochrome c peroxidation by peroxidase. The K(i) for ferricytochrome c is approximately equal to the K(m) for ferrocytochrome c and thus probably accounts for the observed apparent first-order kinetics even at saturating concentrations of ferrocytochrome c. 5. The results are discussed in terms of a possible analogy between the oxidations of cytochrome c catalysed by yeast peroxidase and by mammalian cytochrome oxidase.  相似文献   

17.
The kinetics of reduction and intracomplex electron transfer in electrostatically stabilized and covalently crosslinked complexes between ferredoxin-NADP+ reductase (FNR) and flavodoxin (Fld) from the cyanobacterium Anabaena PCC 7119 were compared using laser flash photolysis. The second-order rate constant for reduction by 5-deazariboflavin semiquinone (dRfH) of FNR within the electrostatically stabilized complex at 10 mM ionic strength (4.0 X 10(8) M-1 s-1) was identical to that for free FNR. This suggests that the FAD cofactor of FNR is not sterically hindered upon complex formation. A lower limit of approximately 7000 s-1 was estimated for the first-order rate constant for intracomplex electron transfer from FNRred to Fldox under these conditions. In contrast, for the covalently crosslinked complex, a smaller second-order rate constant (2.1 X 10(8) M-1 s-1) was obtained for the reduction of FNR by dRfH within the complex, suggesting that some steric hindrance of the FAD cofactor of FNR occurs due to crosslinking. A limiting rate constant of 1000 s-1 for the intracomplex electron transfer reaction was obtained for the covalent complex, which was unaffected by changes in ionic strength. The substantially diminished limiting rate constant, relative to that of the electrostatic complex, may reflect either a suboptimal orientation of the redox cofactors within the covalent complex or a required structural reorganization preceding electron transfer which is not allowed once the proteins have been covalently linked. Thus, although the covalent complex is biochemically competent, it is not a quantitatively precise model for the catalytically relevant intermediate along the reaction pathway.  相似文献   

18.
Phosphate-activated glutaminase in intact pig renal mitochondria was inhibited 50-70% by the sulfhydryl reagents mersalyl and N-ethylmaleimide (0.3-1.0 mM), when assayed at pH 7.4 in the presence of no or low phosphate (10 mM) and glutamine (2 mM). However, sulfhydryl reagents added to intact mitochondria did not inhibit the SH-enzyme beta-hydroxybutyrate dehydrogenase (a marker of the inner face of the inner mitochondrial membrane), but did so upon addition to sonicated mitochondria. This indicates that the sulfhydryl reagents are impermeable to the inner membrane and that regulatory sulfhydryl groups for glutaminase have an external localization here. The inhibition observed when sulfhydryl reagents were added to intact mitochondria could not be attributed to an effect on a phosphate carrier, but evidence was obtained that pig renal mitochondria have also a glutamine transporter, which is inhibited only by mersalyl and not by N-ethylmaleimide. Mersalyl and N-ethylmaleimide showed nondistinguishable effects on the kinetics of glutamine hydrolysis, affecting only the apparent Vmax for glutamine and not the apparent Km calculated from linear Hanes-Woolf plots. Furthermore, both calcium (which activates glutamine hydrolysis), as well as alanine (which has no effect on the hydrolytic rate), inhibited glutamine transport into the mitochondria, indicating that transport of glutamine is not rate-limiting for the glutaminase reaction. Desenzitation to inhibition by mersalyl and N-ethylmaleimide occurred when the assay was performed under optimal conditions for phosphate activated glutaminase (i.e. in the presence of 150 mM phosphate, 20 mM glutamine and at pH 8.6). Desenzitation also occurred when the enzyme was incubated with low concentrations of Triton X-100 which did not affect the rate of glutamine hydrolysis. Following incubation with [14C]glutamine and correction for glutamate in contaminating subcellular particles, the specific activity of [14C]glutamate in the mitochondria was much lower than that of the surrounding incubation medium. This indicates that glutamine-derived glutamate is released from the mitochondria without being mixed with the endogenous pool of glutamate. The results suggest that phosphate-activated glutaminase has a functionally predominant external localization in the inner mitochondrial membrane.  相似文献   

19.
The interaction of three forms of bovine angiotensin-converting enzyme (ACE) with the competitive peptide inhibitor lisinopril with a fluorescent label was studied by the fluorescence polarization technique. The dissociation constants K d of the enzyme-inhibitor complexes in 50 mM Hepes-buffer, pH 7.5, containing 150 mM NaCl and 1 M ZnCl2 at 37°C were (2.3 ± 0.4)·10–8, (2.1 ± 0.3)·10–8, and (2.1 ± 0.2)·10–8 M for two-domain somatic ACE, single-domain testicular ACE, and for the N-domain of the enzyme, respectively. The interaction of the enzyme with the inhibitor strongly depended on the presence of chloride in the medium, and the apparent dissociation constant of the ACE-chloride complex was (1.3 ± 0.2)·10–3 M for the somatic enzyme. The dissociation kinetics of the complex of the inhibitor with somatic ACE did not fit the kinetics of a first-order reaction, but it was approximated by a model of simultaneous dissociation of two complexes with the dissociation rate constants (0.13 ± 0.01) sec–1 and (0.026 ± 0.001) sec–1 that were present at approximately equal initial concentrations. The dissociation kinetics of the single-domain ACE complexes with the inhibitor were apparently first-order, and the dissociation rate constants were similar: (0.055 ± 0.001) and (0.041 ± 0.001) sec–1 for the N-domain and for testicular ACE, respectively.  相似文献   

20.
The effect of luteinizing hormone (luteotropin) and cycloheximide on specific protein synthesis in rat testis Leydig cells has been investigated. Proteins were labelled with either I114C]leucine, [3H]leucine or [35S]methionine during incubation with Leydig-cell suspensions in vitro. Total protein was extracted from the cells and separated by sodium dodecyl sulphate/polyacrylamide-gel electrophoresis. No detectable increase in the synthesis of specific proteins could be observed after incubation of Leydig cells with luteinizing hormone for up to 1 h. However, after a 2h incubation period, an increase in [35S]methionine incorporation was observed in a protein with an apparent mol.wt. of 21000 (referred to as 'protein 21"). When, after labelling of this protein with [35S]-methionine, Leydig cells were incubated for another 30min with cycloheximide, no decrease in radioactivity of this protein band was observed, indicating that it does not have a short half-life. However, another protein band was detected, which after incubation with cycloheximide disappeared rapidly, the reaction following first-order kinetics, with a half-life of about 11 min. This protein, with an apparent mol.wt. of 33000 (referred to as "protein 33"), was found to be located in the particulate fraction of the Leydig cell, and could not be demonstrated in other rat testis-cell types or blood cells. No effect of luteinizing hormone on molecular weight, subcellular localization or half-life of protein 33 was observed. A possible role for protein 33 and protein 21 in the mechanism of action of luteinizing hormone on testosterone production of Leydig cells is discussed.  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