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 46 毫秒
1.
Dispersed acini from dog pancreas were used to examine the ability of dopamine to increase cyclic AMP cellular content and the binding of [3H]dopamine. Cyclic AMP accumulation caused by dopamine was detected at 1·10?8 M and was half-maximal at 7.9±3.4·10?7M. The increase at 1·10?5 M, (7.5-fold) was equal to the half-maximal increase caused by secretin at 1·10?9 M. Haloperidol, a dopaminergic receptor antagonist inhibited cyclic AMP accumulation caused by dopamine. The IC50 value for haloperidol, calculated from the inhibition of cyclic AMP increase caused by 1·10?5 M dopamine was 2.3±0.9·10?6M. Haloperidol did not alter basal or secretin-stimulated cyclic AMP content. [3H]Dopamine binding was studied on the same batch of cells as cyclic AMP accumulation. At 37°C, it was rapid, reversible, saturable and stereospecific. The Kd value for high affinity binding sites was 0.43±0.1·10?7M and 4.7±1.6·10?7M for low affinity binding sites. The concentration of drugs necessary to inhibit specific binding of dopamine by 50% was 1.2±0.4·10/t-7M noradrenaline, 2·10/t-7 M epinine, 4.1±1.8·10/t-6M fluphenazine, 8.0±1.6·10/t-6M haloperidol, 4.2±1.2·10?6Mcis-flupenthixol, 2.7±0.4·10?5Mtrans-flupenthixol, >1·10?5M apomorphine, sulpiride, naloxone and isoproterenol.  相似文献   

2.
Thyroid hormone (T3) has been demonstrated to inhibit the action of aldosterone on sodium transport in toad urinary bladder and rat kidney. We have exammined the effect of T3 on aldosterone action and specific nuclear binding in cultured epithelial cells derived from toad urinary bladder. In cell line TB6-C, addition of 5·10−8 M T3 to culture media for up to 3 days results in no change in short-circuit current or transepithelial resistance. This concentration of T3 completely inhibits the maximal increase in short-circuit current in response to 1·10−7 M aldosterone. The inhibition can be demonstrated with 18 h preincubation or with simultaneous addition of T3 and aldosterone. The half-maximal concentration for the inhibition of the aldosterone effect is approx. 5·10−9 M T3. T3 has no effect on cyclic AMP-stimulated short-circuit current in these cells. The effect of T3 on nuclear binding of [3H]aldosterone was examined using a filtration assay with data analysis by at least-squares curve-fitting program. Best fit was obtained with a model for two binding sites. The dissociation constants for the binding were Kd1 = (0.82 ± 0.36)·10−10 M and Kd2 = (3.2±0.60)·10−8 M.The half-maximal concentration for aldosterone-stimulated sodium transport in these cells is approx. 1·10−8 M. Analysis of nuclear aldosterone binding in cells preincubated for 18 h with 5·10−8 M T3 showed a Kd1 = (0.15 ± 0.10)·10−10 M and Kd2 = (3.5 ± 0.10)·10−8 M. We conclude that T3 i action of aldosterone on sodium transport at a site after receptor binding in the nucleus.  相似文献   

3.
Changes in guanosine cyclic 3′,5′-monophosphate associated with adenosine cyclic 3′,5′-monophosphate and folic acid addition in the presence of ATP have been examined in Dictyostelium discoideum. Preincubation with 1 mM ATP had no effect on the basal cyclic GMP level but increased the cycli GMP accumulation in response to cylci AMP (5·10−8 M) or folic acid (5·10−6 M) 40–50%. ATP could not be replaced by ADP of 5′-adenylyliminodiphosphate. Because ATP has no effect on cyclic AMP receptor binding these results indicate that structural membrane alterations (e.g. membrane phosphorylation) may control the transduction of a chemotactic signal.  相似文献   

4.
Cyclic AMP levels were measured in suspensions of isolated rat liver parenchymal cells during incubation in vitro. Glucagon caused a rapid elevation of cyclic AMP content. With 1.4·10−6 M (5 μg/ml) of the hormone the levels increased about 10-fold during the first minute, thereafter the elevation was less rapid. Maximal values were reached at 5–10 min. Theophylline slightly increased the basal cyclic AMP levels, and markedly augmented the response to glucagon. Teh major part of the cyclic AMP was located within cells, but a siginificant fraction was present in the incubation medium, and the relative amount present extracellularly increased with incubation time. Significant elevation of the cyclic AMP levels was produced by glucagon 1.4·10−10M, and half-maximal stimulation occured at about 2·10−9 M. The initial rate of cyclic AMP accumulation was such more rapid in the parenchymal cells than in liver slices, and the maximal levels obtained were about 3 times higher (comparisons based on the finding that 1 mg liver tissue contains about 105 parenchymal cells). It is concluded that preparations of parenchymal liver cells are useful in the study of glucagon actions on liver tissue.  相似文献   

5.
Amylase secretion and changes in the levels of cyclic AMP and GMP were studied in rabbit parotid gland slices incubated in vitro with a variety of neurohumoral transmitters, their analogs and inhibitors. Cyclic GMP levels increased 8-fold 5 min after exposure to carbachol (10−4 M), without a change in cyclic AMP levels; amylase output also rose. These effects were completely inhibited by muscarinic blockade with atropine, but were unaffected by α-adrenergic blockade with phenoxybenzamine. Epinephrine (4 · 10−5 M) produced a rapid increase in the levels of both cyclic nucleotides and in amylase release. The increase in cyclic GMP level was inhibited by previous exposure of the slices to phenoxybenzamine, while the cyclic AMP rise was prevented by the β-blocking agent, propranolol. Pure α-adrenergic stimulation with methoxamine (4 · 10−4 M) produced modest elevations in cyclic GMP content and amylase output, effects blocked by pre-treatment of slices with either atropine or phenoxybenzamine. At a concentration of 4 · 10−6 M, isoproterenol (a β-agonist) failed to affect cyclic GMP levels, but promptly stimulated increases in cyclic AMP levels, and after a short lag, amylase secretion. At a higher dose (4 · 10−5 M) isoproterenol produced elevations in the levels of both nucleotides. The carbachol-induced effects on cylcic GMP content and amylase release were greatly potentiated by the addition of isoproterenol (4 · 10−6 M).These data strongly suggest that cholinergic muscarinic agonists and α-adrenergic agonist stimulate amylase output in rabbit parotid gland by mechanisms involving cyclic GMP. The atropine-sensitive intracellular events effected by α-stimulation may be dependent upon endogenous generation of acetylcholine. Both cyclic nucleotides seem to be required for the early rapid secretion of amylase. The unique responses achieved by the combination of carbachol and isoproterenol suggest that isoproterenol may increase the sensitivity of this issue to the effects of cholinergic stimuli.  相似文献   

6.
A vasoactive intestinal peptide-sensitive adenylate cyclase in intestinal epithelial cell membranes was characterized. Stimulation of adenylate cyclase activity was a function of vasoactive intestinal peptide concentration over a range of 1 · 10−10−1 · 10−7 M and was increased six-times by a maximally stimulating concentration of vasoactive intestinal peptide. Half-maximal stimulation was observed with 4.1 ± 0.7 nM vasoactive intestinal peptide. Fluoride ion stimulated adenylate cyclase activity to a higher extent than did vasoactive intestinal peptide. Under standard assay conditions, basal, vasoactive inteetinal peptide- and fluoride-stimulated adenylate cyclase activities were proportional to time of incubation up to 15 min and to membrane concentration up to 60 μg protein per assay. The vasoactive intestinal peptide-sensitive enzyme required 5–10 mM Mg2+ and was inhibited by 1 · 10−5 M Ca2+. At sufficiently high concentrations, both ATP (3 mM) and Mg2+ (40 mM) inhibited the enzyme.Secretin also stimulated the adenylate cyclase activity from intestinal epithelial cell membranes but its effectiveness was 1/1000 that of vasoactive intestinal peptide. Prostaglandins E1 and E2 at 1 · 10−5 M induced a two-fold increase of cyclic AMP production. Vasoactive intestinal peptide was the most potent stimulator of adenylate cyclase activity, suggesting an important physiological role of this peptide in the cyclic AMP-dependent regulation of the intestinal epithelial cell function.  相似文献   

7.
The binding and inhibitory properties of 11 benzimidazoles for bovine brain tubulin were investigated. The effects of the benzimidazoles on the initial rates of microtubule polymerization were determined by a turbidimetric assay. The median inhibitory concentrations (I50) for nocodazole, oxibendazole, parbendazole, mebendazole and fenbendazole ranged from 1.97 · 10−6 to 6.32 · 10−6 M. Benomyl, cambendazole and carbendazim had I50 values from 5.83 · 10−5 to 9.01 · 10−5 M. Thiabendazole had an I50 value of 5.49 · 10−4 M. Inhibitor constants (Ki) were determined by the colchicine binding assay. Oxibendazole, fenbendazole, and cambendazole had Ki values of 3.20 · 10−5, 1.73 · 10−5 and 1.10 · 10−4 M, respectively. Oxibendazole and fenbendazole were competitive inhibitors of colchicine. In contrast, cambendazole was a noncompetitive inhibitor of colchicine. The ability of these benzimidazoles to inhibit microtubule polymerization and the mode of action for the anthelmintic benzimidazoles is discussed.  相似文献   

8.
A biotinylated mannotriose (Man3-bio) was dispersively immobilized in the matrix of biotinylated lactose (Gal-Glc-bio) on a streptavidin-covered, 27-MHz quartz crystal microbalance (QCM), and binding kinetics of concanavalin A (Con A) to Man3-bio in the Gal-Glc-bio matrix could be obtained from frequency decreases (mass increases) of the QCM. Association constants (Ka) and binding and dissociation rate constants (kon and koff) could be determined separately as the 1:1 and 1:2 bindings of Con A to Man3-bio on the surface. When Man3-bio was immobilized with content of 1 to 5 mol% in the matrix, the 1:1 binding of Con A to Man3-bio was obtained as Ka = (4 ± 1) × 106 M−1, kon = (4 ± 1) × 104 M−1 s−1, and koff = (12 ± 2) × 10–3 s−1. On the contrary, when Man3-bio was immobilized with content of 20 to 100 mol% in the matrix, the 1:2 binding of Con A to Man3-bio was obtained as Ka = (14 ± 2) × 106 M−1, kon = (14 ± 2) × 104 M−1 s−1, and koff = (7 ± 2) × 10–3 s−1. Thus, Ka for the 1:2 binding was 10 times larger than that for the 1:1 binding, with a three times larger binding rate constant (kon) and a three times smaller dissociation rate constant (koff). This is the first example to obtain separate kinetic parameters for the 1:1 and 1:2 bindings of lectins to carbohydrates on the surface.  相似文献   

9.
Neomycin and related aminoglycosidic antibiotics displace calcium from synaptosomes of guinea pig cerebral cortex and from preparations of phosphatidylinositol diphosphate. At low drug concentrations, inhibition of synaptosomal calcium binding is competitive (Ki = 3·10−5M), at high concentrations it is non-competitive (Ki = 4·10−4M). Monomolecular films of phosphatidylinositol diphosphate are contracted by low concentrations of neomycin in the subphase, and are expanded at high concentrations. This expansion persists even at the collapse pressure indicating a strong interaction between the drug and the lipid.  相似文献   

10.
Vasoactive intestinal peptide (VIP) has been shown to increase cyclic AMP content in isolated epithelial cells of rat ventral prostate. The stimulatory effect of VIP was dependent on time and temperature and was potentiated by a phosphodiesterase inhibitor. At 15°C, the response occurred in the 1·10−10−10−7 M range of VIP concentrations. Half-maximal stimulation of cellular cyclic AMP was obtained at 1.4 nM and maximal stimulation (3-fold basal level) at about 100 nM VIP. Chicken VIP and porcine secretin were agonists of porcine VIP but exhibited a 2-times higher and a 170-times lower potency, respectively. A high concentration (1·10−6 M) of glucagon, somatostatin, neurotensin, substance P, Met-enkephalin or Leu-enkephalin did not modify cAMP levels. The finding of a VIP-stimulated cAMP system in rat prostatic epithelial cells together with the previous characterization of high-affinity receptors for VIP in the same cell preparation, as well as the presence of VIP-containing neurones innervating the male genitourinary tract, strongly suggest that VIP may be involved in prostatic growth regulation and function.  相似文献   

11.
Renoguanylin (REN) is a recently described member of the guanylin family, which was first isolated from eels and is expressed in intestinal and specially kidney tissues. In the present work we evaluate the effects of REN on the mechanisms of hydrogen transport in rat renal tubules by the stationary microperfusion method. We evaluated the effect of 1 μM and 10 μM of renoguanylin (REN) on the reabsorption of bicarbonate in proximal and distal segments and found that there was a significant reduction in bicarbonate reabsorption. In proximal segments, REN promoted a significant effect at both 1 and 10 μM concentrations. Comparing control and REN concentration of 1 μM, JHCO3, nmol cm− 2 s− 1 − 1,76 ± 0,11control × 1,29 ± 0,08REN 10 μM; P < 0.05, was obtained. In distal segments the effect of both concentrations of REN was also effective, being significant e.g. at a concentration of 1 μM (JHCO3, nmol cm− 2 s− 1 − 0.80 ± 0.07control × 0.60 ± 0.06REN 1 μM; P < 0.05), although at a lower level than in the proximal tubule. Our results suggest that the action of REN on hydrogen transport involves the inhibition of Na+/H+exchanger and H+-ATPase in the luminal membrane of the perfused tubules by a PKG dependent pathway.  相似文献   

12.
Two processes of porphyrins in heterogeneous systems containing aqueous and membrane phases have been studied with hematoporphyrin and hematoporphyrin derivative: Dimerization equilibrium in the aqueous phases and porphyrin-membrane binding equilibrium using liposomes as models for biological membranes. The interrelationship of aqueous aggregations and membrane binding was probed and the porphyrin aggregation state in the membrane, at equilibrium, was assessed. Fluorimetric techniques were employed. The dimerization equilibrium constants, at neutral pH and 37°C were found to be 2.8 · 105 M−1 and 1.9 · 106 M−1 for hematoporphyrin and its derivative, respectively. Over a porphyrin concentration range going from monomer-dominant to dimer-dominant systems, we have found that only monomers are bound to the membrane. The respective monomer-liposome binding constants, found to be independent of the initial monomer/dimer distribution in the aqueous phase, were determined to be 1.6 · 103 M−1 and 4.1 · 103 M−1 at neutral pH and 37°C for hematoporphyrin and its derivative, respectively. The monomer-liposome interaction was found to perurb the initial monomer/dimer distribution in the aqueous phase, so that the monomers residing at equilibrium in the membrane originate from both monomers and dimers in the aqueous phase.  相似文献   

13.
Synthetic peptide S3Δ has high affinity for bacterial endotoxin or lipopolysaccharide (LPS). Under tested conditions of pH 5–9 and 0–0.4 M NaCl, the affinity constant, KD ranged from 2·10−6 to 2·10−9 M−1. A novel affinity matrix based on peptide S3Δ was developed for removal of LPS from solutions such as: water; buffers with a wide range of ionic strength and pH; medium for cell culture; and protein solutions under optimized conditions. At a starting LPS of ≈100 EU/ml, a post-purification level below 0.005 EU/ml was achieved.  相似文献   

14.
Galactomannans isolated from legume seed endosperms, including those of commercial interest, have been characterized by multidetection aqueous SEC. Galactomannans derived from seeds of the Faboideae subfamily had substantially higher Mw than those from Caesalpinioideae seeds (Mw,Fab = 2.4–3.1 × 106 g/mol, Mw,Caes. = 0.86–2.1 × 106 g/mol) and within the latter botanical subfamily, an apparent correlation between Mw and the degree of galactose substitution DG was found. The molar mass distributions were unimodal and differed primarily by a scale factor, with distributional widths narrower than a true Flory ‘most-probable distribution’; good fits to Schulz–Zimm model were obtained. Across subfamilies no differences were found in the exponents of [η]–M and RvM relationships (0.61 ± 0.02, 0.54 ± 0.01, respectively), the Flory chain stiffness ratio (C = 20 ± 1 (BSF analysis)), or the persistence length (Lp = 5.5 ± 0.2 nm) obtained from SEC fraction data. However, it was found that prefactors in the [η]–M and RvM relationships as well as the unperturbed parameter KΘ decrease in proportion to DG and therefore chain density. Generalized relationships incorporating galactose-dependent prefactors were therefore developed to model SEC fraction data of native galactomannans ([η]GM = (1800 ± 200) × Mo−1.61 × M0.61±0.02, Rv,GM = 0.63 ± 0.05 × Mo−0.54 × M0.54±0.01) as well as lower-M fractions obtained by ultrasonication ([η]GM = (730 ± 100) × Mo−1.71 × Mw0.71±0.02, Rv,GM = 0.49 ± 0.05 × Mo−0.57 × Mw0.57±0.01, M ≈ 1 × 105-native). As a consequence of this dependence and the observed patterns in molar mass variation, [η] varies within a narrow range for galactomannans as a whole despite substantial Mw differences.  相似文献   

15.
The existence of two types of binding sites for ouabain in human erythrocyte membranes is described. Receptor sites designated as ‘type I’, which may be identical to the K+-insensitive sites of intact cells, were detected at concentrations of ouabain as low as 10−7 M. The ‘type II’ receptor sites require the inclusion of Mg2+ + Pi to form complexes with ouabain; they may be identical to the K+-sensitive sites of intact cells. These sites were saturated at approx. 5 · 10−7 M ouabain but could not be detected at higher concentrations. The range of ouabain concentrations at which ‘type I’ receptors start to predominate (i.e. 5 · 10−8–5 · 10−7 M) was termed ‘critical digitalis concentrations’. The process of binding reached equilibrium within 1 and 4 h for ‘type I’ and ‘type II’ sites, respectively. The dissociation constant for ‘type II’ receptor-ouabain complexes was 7.6 · 10−9 M.Under similar experimental conditions, rat erythrocyte membranes exhibited only non-saturable sites.Alterations in the proportions of the two types of receptors were demonstrated by preincubation of the membranes, in the presence or absence of Mg2+ + Pi, prior to the addition of ouabain. In the first case, ‘type II receptor-ouabain’ complexes were stabilized at about 50% of the untreated membranes and ‘type I-ouabain’ complexes slowly approached equilibrium over a period of 24 h. In the latter instance, ‘type I’ receptors were not detected, and only ‘type II-ouabain’ complexes prevailed.  相似文献   

16.
In this study, the hydraulic conductivity (Lp), Me2SO permeability ( Me2SO), and the reflection coefficients (ς) and their activation energies were determined for Metaphase II (MII) mouse oocytes by exposing them to 1.5 M Me2SO at temperatures of 30, 20, 10, 3, 0, and −3°C. These data were then used to calculate the intracellular concentration of Me2SO at given temperatures. Individual oocytes were immobilized using a holding pipette in 5 μl of an isosmotic PBS solution and perfused with precooled or prewarmed 1.5 M Me2SO solutions. Oocyte images were video recorded. The cell volume changes were calculated from the measurement of the diameter of the oocytes, assuming a spherical shape. The initial volume of the oocytes in the isoosmotic solution was considered 100%, and relative changes in the volume of the oocytes after exposure to the Me2SO were plotted against time. Mean (means ± SEM) Lpvalues in the presence of Me2SO ( Me2SOp) at 30, 20, 10, 3, 0, and −3°C were determined to be 1.07 ± 0.03, 0.40 ± 0.02, 0.18 ± 0.01, 7.60 × 10−2± 0.60 × 10−2, 5.29 × 10−2± 0.40 × 10−2, and 3.69 × 10−2± 0.30 × 10−2μm/min/atm, respectively. The Me2SOvalues were 3.69 × 10−3± 0.3 × 10−3, 1.07 × 10−3± 0.1 × 10−3, 2.75 × 10−4± 0.15 × 10−4, 7.83 × 10−5± 0.50 × 10−5, 5.24 × 10−5± 0.50 × 10−5, and 3.69 × 10−5± 0.40 × 10−5cm/min, respectively. The ς values were 0.70 ± 0.03, 0.77 ± 0.04, 0.81 ± 0.06, 0.91 ± 0.05, 0.97 ± 0.03, and 1 ± 0.04, respectively. The estimated activation energies (Ea) for Me2SOp, Me2SO, and ς were 16.39, 23.24, and −1.75 Kcal/mol, respectively. These data may provide the fundamental basis for the development of more optimal cryopreservation protocols for MII mouse oocytes.  相似文献   

17.
Dispersed acini from dog pancreas were used to examine the ability of dopamine to increase cyclic AMP cellular content and the binding of [3H] dopamine. Cyclic AMP accumulation caused by dopamine was detected at 1.10(-8) M and was half-maximal at 7.9 +/- 3.4 10(-7) M. The increase at 1.10(-5) M, (7.5-fold) was equal to the half-maximal increase caused by secretin at 1.10(-9) M. Haloperidol, a dopaminergic receptor antagonist inhibited cyclic AMP accumulation caused by dopamine. The IC50 value for haloperidol, calculated from the inhibition of cyclic AMP increase caused by 1.10(-5) M dopamine was 2.3 +/- 0.9.10(-6) M. Haloperidol did not alter basal or secretin-stimulated cyclic AMP content. [3H] Dopamine binding was studied on the same batch of cells as cyclic AMP accumulation. At 37 degrees C, it was rapid, reversible, saturable and stereospecific. The Kd value for high affinity binding sites was 0.43 +/- 0.1.10(-7) M and 4.7 +/- 1.6.10(-7) M for low affinity binding sites. The concentration of drugs necessary to inhibit specific binding of dopamine by 50% was 1.2 +/- 0.4.10(-7) M epinine, 4.1 +/- 1.8.10(-6) M fluphenazine, 8.0 +/- 1.6.10(-6) M haloperidol, 4.2 +/- 1.2.10(-6) M cis-flupenthixol, 2.7 +/- 4.0.10(-5) M trans-flupenthixol, less than 1.10(-5) M apomorphine, sulpiride, naloxone and isoproterenol.  相似文献   

18.
The effect of the local anesthetic dibucaine on the solid to liquid-crystalline phase transition in phospholipid vesicles was studied by calorimetry and fluorescence polarization. The partition coefficient (> 3000) of dibucaine in the membranes of vesicles prepared from acidic phospholipids was more than 20 times higher than in neutral phospholipid membranes under the same conditions. Calorimetric measurements on vesicles prepared form acidic phospholipids (bovine brain phosphatidylserine; dipalmitoylphosphatidylglycerol) showed that dibucaine (1 · 10−4M) produced a significant reduction in the gel-liquid crystalline transition temperature (Tc). This fluidizing effect of dibucaine on acidic phospholipid membranes was even more marked in the presence of Ca2+. In contrast, dibucaine at the same concentration did not alter the Tc of neutral phospholipids (dipalmitoylphosphatidylcholine). Significant increase in the fluidity of neutral phospholipid membranes occurred only at higher dibucaine concentrations (2 · 10−3M. Measurements of the fluorescence polarization and lifetime of the probe, 1,6-diphenylhexatriene, in acidic phospholipid vesicles revealed that dibucaine (1 · 10−4M caused an increase in the probe rotation rate indicating an increase in the fluidity of the phospholipid membranes. A good correlation was obtained between fluorescence polarization data on dibucaine-induced changes in membrane fluidity and calorimetric measurements on vesicles of the same type.  相似文献   

19.
The effects of various 11-deoxyprostaglandin E analogs on the basal and prostaglandin E2 (PGE2)-induced cyclic AMP accumulation in the rat anterior pitutiary were studied . 13-Hydroxy-9-oxoprost-14-ynoic acid at 5 × 10−4M, but not 5 × 10−5M, decreased (45%) the induced accumulation and did not alter the basal accumulation; 15-hydroxy-9-oxoprost-13-ynoic acid at 5 × 10−4M caused less of a decrease (29%) in the induced and also did not alter the basal accumulation. (14Z)-13-Hydroxy-9-oxoprost-14-enoic acid at 5 × 10−4M did not alter the induced and caused a slight increase (5 fold) in the basal accumulation. 7-Oxa-13-prostynoic acid increased slightly the basal accumulation at 5 × 10−5M (2 fold) and 2.33 × 10−4M (6 fold) and did not antagonize the induced accumulation. Thus, the 9-ketoprostynoic acids are effective PGE2 antagonists in this system.  相似文献   

20.
The catalytic subunit of cAMP-dependent protein kinase from rat adipose tissue was purified to apparent homogeneity by making use of the differential binding of the holoenzyme and the free catalytic subunit to CM-Sephadex and by gel chromatography. Stability and yield was improved by inclusion of nonionic detergent in all steps after dissociation of the holoenzyme. Isoelectric focusing separated enzyme species with pI values of 7.8 and 8.6–8.8. The amino acid composition was similar to the enzyme purified from other tissues. Enzyme activity was markedly unstable in dilute solutions (<5 μg/ml). Additions of nonionic detergent, glycerol, bovine serum albumin and, especially, histones stabilized the enzyme. With protamine, the catalytic subunit had an apparent Km of 60 μM and Vmax of 20 μmol·min−1·mg−1, corresponding values with mixed histones were 12 μM and 1.2 μmol·min−1·mg−1. With both protein substrates the apparent Km for ATP was 11 μM. Concentrations of Mg2+ above 10 mM were inhibitory. Histone phosphorylation was inhibited by NaCl (50% at 0.5 M NaCl) while protamine phosphorylation was stimulated (4-fold at 1 M NaCl). Inorganic phosphate inhibited both substrates (histones: 50% at 0.3 M, and protamine: 50% at 0.5 M). pH optimum was around pH 9 with both substrates. The catalytic subunit contained 2.0 (range of three determinations, 1.7–2.3) mol phosphate/mol protein. It was autophosphorylated and incorporated 32Pi from [γ-32P]ATP in a time-dependent process, reaching saturation when approx. 0.1 mol phosphate/mol catalytic subunit was incorporated.  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