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 30 毫秒
1.
—The intracellular disposition of the convulsant agent, methionine sulphoximine (MSO), administered as methyl-labelled [3H]MSO, was examined in rat brain. Intraperitoneal (i.p.) and intrathecal (i.th.) routes were compared. The effect of simultaneous administration of methionine on the uptake, the regional distribution and the intracellular disposition of [3H]MSO was also assessed: (1) The peak uptake of i.p. [3H]MSO was at 2 h and amounted to about 1 per cent of the dose; the peak uptake of i.th. [3H]MSO was at 30 min post-injection and amounted to 40 per cent of the administered dose. The uptake was effectively reduced when methionine was simultaneously administered. (2) The regional distribution of [3H]MSO as a function of time after injection revealed a rather uniform penetration of the entire brain by the drug. A maximum of 43 per cent of the tissue radioactivity was found in the cerebellum 2 h after i.p. injection, while 49 per cent accumulated in the extracortical portion of the brain 3·5 h after i.th. administration. Methionine did not affect the regional distribution of [3H]MSO. (3) Differential centrifugation of samples of cortex and cerebellum revealed an association of [3H]MSO with intracellular particulate fractions. Since closely similar proportions of MSO occurred in the crude mitochondrial and the microsomal fractions, these fractions were analysed further: (a) [3H]MSO was bound to nerve endings sedimenting at the 1·0 m–1·2 m-sucrose interface; this binding was not abolished by prior increase of the endogenous cerebral methionine pool; and (b) [3H]MSO was released by subjecting the nerve endings to osmotic shock. However, the striking finding was that [3H]MSO could not be released from the nerve endings of the cerebellum from animals pre-treated with methionine. (4) An association of [3H]MSO was observed with the membranes of the endoplasmic reticulum and specifically with its agranular component. (5)The results implicate the cerebellum as the primary target for MSO, in confirmation of the original observations of Lodin (1958).  相似文献   

2.
Angiosperms possess a retaining trans-α-xylosidase activity that catalyses the inter-molecular transfer of xylose residues between xyloglucan structures. To identify the linkage of the newly transferred α-xylose residue, we used [Xyl-3H]XXXG (xyloglucan heptasaccharide) as donor substrate and reductively-aminated xyloglucan oligosaccharides (XGO–NH2) as acceptor. Asparagus officinalis enzyme extracts generated cationic radioactive products ([3H]Xyl·XGO–NH2) that were Driselase-digestible to a neutral trisaccharide containing an α-[3H]xylose residue. After borohydride reduction, the trimer exhibited high molybdate-affinity, indicating xylobiosyl-(1→6)-glucitol rather than a di-xylosylated glucitol. Thus the trans-α-xylosidase had grafted an additional α-[3H]xylose residue onto the xylose of an isoprimeverose unit. The trisaccharide was rapidly acetolysed to an α-[3H]xylobiose, confirming the presence of an acetolysis-labile (1→6)-bond. The α-[3H]xylobiitol formed by reduction of this α-[3H]xylobiose had low molybdate-affinity, indicating a (1→2) or (1→4) linkage. In NaOH, the α-[3H]xylobiose underwent alkaline peeling at the moderate rate characteristic of a (1→4)-disaccharide. Finally, we synthesised eight non-radioactive xylobioses [α and β; (1↔1), (1→2), (1→3) and (1→4)] and found that the [3H]xylobiose co-chromatographed only with (1→4)-α-xylobiose. We conclude that Asparagus trans-α-xylosidase activity generates a novel xyloglucan building block, α-d-Xylp-(1→4)-α-d-Xylp-(1→6)-d-Glc (abbreviation: ‘V’). Modifying xyloglucan structures in this way may alter oligosaccharin activities, or change their suitability as acceptor substrates for xyloglucan endotransglucosylase (XET) activity.  相似文献   

3.
Three distinct nuclease activities, degrading double-stranded substrates, were isolated from the ribosomal salt wash fraction of Ehrlich ascites tumor cells. One of them is an absolutely Mn2+-dependent RNase H, capable of degrading the polyribonucleotide strand of a poly(A) · poly(dT) hybrid only. The other two nuclease activities are: a Mg2+-dependent RNase H and a Mn2+-dependent ribonuclease, specific for double-stranded RNA. These two activities were inseparable by DEAE-cellulose and phosphocellulose chromatography and both were completely inhibited by 20 mmN-ethymaleimide. It is possible that one protein molecule is responsible for the two activities, depending on the nature of the metal ion, though the existence of two different enzyme molecules is not excluded. The three activities are most probably of extranucleolar origin. A function for the double-stranded RNA-specific enzyme is suggested in the processes regulating protein synthesis. The role of the RNase H activities isolated from the ribosomal salt wash fraction is unclear.  相似文献   

4.
[3H]norepinephrine was shown to bind to specific sites on isolated fat cells. A Scatchard plot of norepinephrine binding showed two apparent Ka of 1.9 · 106 and 1.2 · 105 LM?. 1.4 · 10?4 M Norepinephrine covalently-linked to agarose beads reduced [3H]norepinephrine binding by over 50%. Several structurally related drugs were compared as inhibitors of [3H]norepinephrine binding and as stimulators of lipolysis in preparations of similarly prepared cells. Dose-response curves for norepinephrine, epinephrine and isoproterenol showed the affinities for binding inhibition and for stimulation of lipolysis to be in the same range of 6 · 10?7-2 · 10?6 M. Dopamine and dopa were potent inhibitors of [3H]norepinephrine binding at 8.5 · 10?7 M and 2.0 · 10?6 M respectively, but did not stimulate lipolysis even at 10?4 M. Propranolol, a β-adrenergic antagonist, had no effect on [3H]norepinephrine binding at 10?4 M but completely inhibited catecholamine-stimulated lipolysis at 10?5 M. Phentolamine, an α-adrenergic antagonist, did not inhibit binding or catecholamine-stimulated lipolysis at 10?4 M. Ephedrine, metaraminol, phenylephrine and normetanephrine were also ineffective both as [3H]norepinephrine binding inhibitors and as stimulators of lipolysis. The results suggested the catechol ring of catecholamines is more important than the ethanolamine side chain as a requirement for binding, while both an intact catechol moiety and ethanolamine function appear necessary for physiological effect.  相似文献   

5.
The binding characteristics of the β-adrenergic agonist (±)-[3H]hydroxybenzylisoproterenol to rat adipocyte membranes were studied. Binding was rapid, reaching equilibrium within 10 min at 37°C (second order rate constant k1=1.37·107·M?1·min?1). Dissociation of specific binding by 0.5 mM (?)-isoproterenol suggested dissociation from two different sites with respective dissociation rate constants k2 of 0.106·min?1 and 0.011·min?1.[3H]Hydroxybenzylisoproterenol binding was saturable (Bmax=690±107 fmol/mg protein), yielding curvilinear Scatchard plots. Computer modeling of these data were consistent with the existence of two classes of [3H]hydroxybenzylisoproterenol binding sites, one having high affinity (KD=3.5±0.7 nM) but low binding capacity (10% of the total sites) and one haveing low affinity (KD=101±20 nM) but high binding capacity (90% of the sites). Adrenergic ligands competed with [3H]hydroxybenzylisoproterenol binding with the following order of potency=(?)-propranolol>(?)-isoproterenol>(?)-norepinephrine≈ (?)-epinephrine>>(+)-isoproterenol=(+)-propranolo, which is consistent with binding to β1-adrenergic receptors. Competition curves of [3H]hydroxybenzylisoproterenol binding by the β-agonist (?)-isoproterenol were shallow and modeled to two affinity states of binding, whereas, competition curves by β-antagonist (?)-propranolol were steeper with Hill number near to one. Gpp[NH]p severely reduced [3H]hydroxybenzyl-isoproterenol binding, an effect which apparently resulted from the reduction of the number of both the high and low affinity sites. In membranes which had been previously exposed to (?)-isoproterenol, then number of [3H]hydroxybenzylisoproterenol binding sites was reduced by 50%, an effect which apparently resulted from the loss of part of both the high and low affinity state binding sites. Finally, the ability of (?)-isoproterenol to stimulate adenylate cyclase correlate closely with the ability of (?)-isoproterenol to displace [3H]hydroxybenzylisoproterenol binding. Comparison of these findings with the binding characteristics of the β-antagonist [3H]dihydroalprenolol to rat adipocyte membranes, led to conclude that [3H]hydroxybenzylisoproterenol can be successfully used to label the β-adrenergic receptors of rat fat cells and suggests that it might be a better ligand than [3H]dihydroalprenolol in these cells.  相似文献   

6.
[3H]Prostaglandin D2 binding to rabbit platelets was increased by about 150% in the presence of β-adrenoceptor agonist, isoproterenol. The isoproterenol-induced potentiation of the [3H]prostaglandin D2 binding gave a bell-shaped dose-response relationship (maximum response at 3·10−8 M) in a stereospecific manner. Similar and moderate potentiation was obtained with terbutaline. On the other hand, β-adrenoceptor antagonists such as alprenolol, propranolol and butoxamine (β2-specific) had no potentiating effect on [3H]prostaglandin D2 binding; rather, they abolished the isoproterenol-induced increase of [3H]prostaglandin D2 binding. The β1-specific antagonist, metoprolol, did not have any effect. Rabbit platelets were found to possess one [3H]prostaglandin D2 binding site (Kd = 6·10−7 M, Bmax = 787 fmol/mg protein). In the presence of isoproterenol at 3·10−8 M, Bmax was increased with unaltering Kd value. Isoproterenol did not increase [3H]prostaglandin E1, [3H]prostaglandin E2 and [3H]prostaglandin F bindings to platelets. The potential effect of isoproterenol was mimicked by forskolin, theophylline, dibutyryl cyclic AMP, prostaglandin E1 and prostaglandin I2, but it was abolished by 2′, 5′-dideoxyadenosine, an inhibitor of adenylate cyclase, indicating that elevated level of cyclic AMP may be available for the induction of the increase of [3H]prostaglandin D2 binding. Prostaglandin D2-induced cyclic AMP synthesis and antiaggregation activity were also augmented in the presence of isoproterenol. These results suggest a β2-adrenoceptor-mediated cyclic AMP-dependent mechanism for the regulation of prostaglandin D2 receptor binding in rabbit platelets.  相似文献   

7.
A series of metal complexes of La(III) and Th(IV) have been synthesized with newly derived biologically active ligands. These ligands were synthesized by the condensation of 3-substituted-4-amino-5-hydrazino-1,2,4-triazole with 8-formyl-7-hydroxy- 4-methylcoumarin. The structure of the complexes has been proposed by elemental analyses, spectroscopic data i.e. i.r., 1H nmr, Uv-Vis, FAB-mass and thermal studies. The elemental analyses of the complexes conform to the stoichiometry of the type [La(L)·3H2O]·2H2O and [Th(L)(NO3)·2H2O]·2H2O where (L = LI-LIV). All the complexes are soluble in DMF and DMSO and are non-electrolytes in DMF and DMSO. All these ligands and their complexes have also been screened for their antibacterial (Escherichia coli, Staphylococcus aureus, Staphylococcus pyogenes and Pseudomonas aeruginosa) and antifungal activities (Aspergillus niger, Aspergillus flavus and cladosporium) by the MIC method. The brine shrimp bioassay was also carried out to study their invitro cytotoxic properties.  相似文献   

8.
Cultured chick embryo cardiac myoblasts specifically bind [3H]nonrepinephrine. The binding is rapid and reversible. Bound [3H]nonrepinephrine, dissociated by 1 M HCl, can be rebound to fresh cells. β-Adrenergic catecholamines were most potent in displacing [3H]nonrepinephrine from the cellular bindign sites. The binding reaction did not show stereospecificity. α-Adrenergic amines were much less potent. Propranolol, but no phentolamine, competed for the sites. Approximately 2.5 · 106 specific binding sites are present per myocardial cell. The sites appear to be present predominantly at the cell surface in that nonrepinephrine linked to agarose beads competes for th sites. Similarly, the sites were degraded by either trypsin or trypsin bound to agarose. Two different binding constants, K = 2 · 106 and 1 · 105, were observed. Proteolytic enzymes decreased binding whereas certain hospholipases led to an increase in specific binding. Divalent cations at concentrations > 1 mM diminished binding as did chelating agents.  相似文献   

9.
A double-stranded RNA specific nuclease (ds RNase) has been purified from the pearl milletPennisetum typhoides. The purification involved S-30 preparation from the germinating embryos, DEAE-cellulose and DNA-cellulose chromatography. The partially pure enzyme preferentially solubilized the synthetic double-stranded polynucleotide [3H]poly(rA) · poly(rU); the degradation of [3H]poly(rC) was fourteen fold lower under the same assay conditions. Further more, the ds RNase activity was inhibited to an extent of 58% by ethidium bromide, which is known to intercalate with double-stranded RNAs. Active sulfhydryl groups were found to be necessary for the ds RNase activity since the enzyme action was inhibited by N-ethylmaleimide. Ethidium bromide and N-ethyl-maleimide did not significantly inhibit the ss RNase activity. In contrast, diethyl pyrocarbonate inhibited ss RNase activity completely and ds RNase by 58%. Heating the enzyme for 20 min at 50°C resulted in drastic loss of both enzyme activities. The ds RNase showed maximum activity in the pH range of 6.5 to 7.5. The enzyme actsin vitro onE. coli 30S precursor ribosomal RNA and the cleavage products migrated in the region of mature 23S and 16S rRNAs.  相似文献   

10.
[3H]Dihydroalprenolol, a potent ß-adrenergic antagonist, was used to identify the adenylate cyclase-coupled ß-adrenoceptors in isolated membranes of rat skeletal muscle. The receptor sites, as revealed [3H]dihydroalprenolol binding, were predominantly localized in plasmalemmal fraction. That skeletal muscle fraction may also contain the plasmalemma of other intramuscular cells, especially that of blood vessels. Hence, the [3H]dihydroalprenolol binding observed in that fraction may be due partly to its binding to the plasmalemma of blood vessels. Small but consistent binding was also observed in sarcoplasmic reticulum and mitochondria. The level of [3H]dihydroalprenolol binding in different subcellular fractions closely correlated with the level of adenylate cyclase present in those fractions.The binding of [3H]dihydroalprenolol to plasmalemma exhibited saturation kinetics. The binding was rapid, reaching equilibrium within 5 min, and it was readily dissociable. From the kinetics of binding, association (K1) and dissociation (K2) rate constants of 2.21 · M? · min?1 and 3.21 · 10?1, respectively, were obtained. The dissociation constant (Kd) of 15 nM for [3H]dihydroalprenolol obtained from saturation binding data closely agreed with the (Kd) derived from the ratio of dissociation and association rate constants (K2/K1).Several β-adrenergic agents known to be active on intact skeletal muscle also competed for [3H]dihydroalprenolol binding sites in isolated plasmalemma with essentially similar selectivity and stereospecificity. Catecholamines competed for [3H]dihydroalprenolol binding sites with a potency of isoproterenol > epinephrine > norepinephrine. A similar order of potency was noted for catecholamines in the activation of adenylate cyclase. Effects of catecholamines were stereospecific, (?)-isomers being more than potent than (+)-isomers. Phenylephrine, an α-adrenergic agonist, showed no effect either on [3H]dihydroalprenolol binding or on adenylate cyclase. Known ß-adrenergic antagonists, propranolol and alprenolol, stereospecifically inhibited the [3H]dihydroalprenolol binding and the isoproterenol-stimulated adenylate cyclase. The (Ki) values for the antagonists determined from inhibition of [3H]dihydroalprenolol binding agreed closely with the (Ki) values obtained from the inhibition of adenylate cyclase. The data suggest that the binding of [3H]dihydroalprenolol in skeletal muscle membranes possess the characteristics of a substance binding to the ß-adrenergic receptor.  相似文献   

11.
The kinetic properties and subcellular distribution of an esterifying enzyme in the pigment epithelium of bovine retina have been studied using both [1-3H]retinol and [3H]retinol bound to cellular retinol-binding protein as substrates. The most active esterifying fraction in pigment epithelial cell preparations was the microsomes, but the lysosome plus mitochondria fraction also showed some activity, probably due to endoplasmic reticulum present as an impurity. The microsomal enzyme showed optimum activity at pH 7.5, and the reaction was linear up to 30 μg protein and for the first 10–15 min. The apparent Km values were 16.6 · 10?6 and 5.5 · 10?6 M for [3H]retinol and bound [3H]retinol, respectively. This is the first time that retinol bound to cellular retinol-binding protein has been shown to undergo metabolic stransformation. The microsomal esterifying activity was destroyed by boiling for 1 min, or after freezing for 2 months. No clear requirement for ATP, CoA or fatty acid could be demonstrated.Of all the other tissues examined under the same experimental conditions as those used for the pigment epithelium, onlt intestine showed measurable activity. With larger amounts of tissue protein and longer incubation periods, activity was also detectable in microsomes of liver, testis and retina  相似文献   

12.
Reaction of gibberellin A3 (GA3) with carrier-free tritium gas and 5% palladium on calcium carbonate as catalyst gave a complex mixture of products, several of which were isolated and identified. Three of the purified products are the radioactive forms of naturally occurring gibberellins: [3H]GA3 (1), [3H]GA1 (2) and [3H]tetrahydro GA3 (4). Another substance was isolated and tentatively identified as [3H]16,17-dihydro GA3 (3). GLC was used to determine the specific activities of 1 and 2. [3H]GA3 likely arises from palladium catalysed nonspecific exchange of GA3 alkane hydrogen atoms with tritium. [3H]GA1 is also exchange labeled but most of its radioactivity is due to tritium addition to the C-1,2 olefinic bond of GA3.  相似文献   

13.
—Cortex slices of rat brain were incubated with glucose mixed-labelled with 3H and 14C in the 6-position and the 3H/14C ratios of lactate, acetate, citrate and acetylcholine were determined. The values obtained were: lactate 0·95, acetate 0·85, citrate 0·65 and acetylcholine 0·67 when expressed in relation to a glucose 3H/14C ratio of 1·00. When brain slices were incubated with [2-14C, 2-3H]acetate in the presence of unlabelled glucose, labelled acetylcholine was formed with a 3H/14C ratio not significantly different from the labelled substrate. The results indicate that citrate is a precursor to the acetyl moiety of acetylcholine.  相似文献   

14.

Aims

The purpose of this study was to clarify the expression and function of peptide transporter 2 (PEPT2) in primary cultured alveolar type II epithelial cells and in transdifferentiated type I-like cells.

Main methods

Real-time PCR analysis, uptake study of [3H]Gly-Sar, and immunostaining were performed in alveolar epithelial cells.

Key findings

The expression of PEPT2 mRNA in type II cells isolated from rat lungs was highest at day 0, and decreased rapidly during culture of the cells. In accordance with this change, PEPT2 activity estimated as cefadroxil-sensitive [3H]Gly-Sar uptake also decreased along with transdifferentiation. The expression of PEPT2 protein in type II cells was confirmed by immunostaining and Western blot analysis. The uptake of [3H]Gly-Sar in type II cells was time- and pH-dependent. In contrast, minimal time-dependence and no pH-dependence of [3H]Gly-Sar uptake were observed in type I-like cells. The maximal [3H]Gly-Sar uptake was observed at pH 6.0, and the uptake decreased at higher pHs in type II cells. The uptake of [3H]Gly-Sar in type II cells was inhibited by cefadroxil in a concentration-dependent manner, the IC50 value being 4.3 μM. On the other hand, no significant inhibition by cefadroxil was observed in type I-like cells. In addition, [3H]Gly-Sar uptake in type II cells was saturable, the Km value being 72.0 μM.

Significance

PEPT2 is functionally expressed in alveolar type II epithelial cells, but the expression decreases along with transdifferentiation, and PEPT2 would be almost completely lost in type I cells.  相似文献   

15.
Polysomes were isolated from Aspergillus niger and were characterized on sucrose gradients in several ways. First, they were found to be susceptible to degradation by treatment with RNase or EDTA. Second, they were labeled after treating mycelia with short pulses of [3H]uridine or [3H]leucine prior to polysome isolation. Third, they were capable of stimulating incorporation of [3H]leucine into trichloroacetic acid-precipitable material in a chick reticulocyte cell-free protein-synthesizing system. When isolated [3H]leucine pulse-labeled polysomes were treated with either EDTA-RNase or puromycin, 80–90% of the radioactivity was released, indicating that only the nascent polypeptide chains were labeled. After exposing mycelia for 1 min to [14C]mannose, the polysomes were exclusively labeled, indicating that initial glycosylation takes place on nascent polypeptide chains. Preincubation of mycelia with 2-deoxyglucose followed by pulse-labeling with [3H]leucine and [14C]mannose showed that 2-deoxy-d-glucose inhibits both protein synthesis and glycosylation. However, similar preincubation with tunicamycin caused an 80% drop in [14C]mannose label in the polysomes, but only a 10–20% drop of [3H]leucine label, suggesting that glycosylation of nascent chains in A. niger involves an oligosaccharide-lipid intermediate, since it has been shown that tunicamycin inhibits the synthesis of such an intermediate. When isolated polysomes were placed into an in vitro glycosylating mixture containing Mn2+, GDP-[14C]mannose, and smooth membranes from A. niger nascent chains were labeled. This reaction was shown to be dependent on addition of polysomes to the mixture and was not inhibited by 2-deoxy-d-glucose or tunicamycin. Both in vivo and in vitro glycosylated nascent chains were found to have about the same size range, and so it is suggested that in vitro no new oligosaccharide chains were synthesized, but preexisting chains were extended.  相似文献   

16.
Abstract— A new combined ion-exchange and thin-layer-chromatographic procedure is described which separates and measures quantitatively, after intraventricular injection of [3H]dopamine (DA), the rat brain content of labelled noradrenaline (NA) and the following labelled noradrenaline metabolites: free 3-methoxy-4-hydroxyphenylethyleneglycol (MOPEG), conjugated MOPEG, free plus conjugated dihydroxyphenylethyleneglycol (DOPEG), vanillic mandelic acid (VMA) and normetanephrine (NM). Labelled dopamine and its metabolites were also measured. The time-course study performed from 5 min to 24 h after [3H]DA showed that MOPEG and DOPEG, mainly as conjugates, are major NA metabolites whereas VMA is a very insignificant NA metabolite in the rat brain. A very rapid initial increase of [3H]NM, free MOPEG and conjugated MOPEG was found during the time interval where the [3H]NA biosynthesis is very high (0–15 min). This combined with the finding that these metabolites stabilize at lower levels during the [3H]NA ‘storage phase’ (9–24 h) provides a strong indication that newly synthesized NA preferentially is metabolized. Our measurements of endogenous NA, free MOPEG and conjugated MOPEG provide additional support. The injections of various decreasing doses of [3H]DA (3·08–0·0010 μg) showed that the proportions of total [3H]MOPEG and total [3H]DOPEG to [3H]NA were constant after all [3H]DA doses investigated. This finding indicates that the [3H]NA synthesized in situ behaves as a tracer, even after injections of non-tracer doses of [3H]DA. The results seem thus to indicate that the present technique provides a powerful tool for the investigations on central noradrenaline metabolism.  相似文献   

17.
Hydroxylation of gibberellin-[3H] A1 (GA1-[3H]) to GA8-[3H] by the 95000 g supernatant fluid from imbibed bean seeds required Fe2+ or Fe3+ and O2 but was insensitive to CO. The hydroxylating enzyme has a sedimentation coefficient of 4·5 S, and was precipitated by (NH4)2SO4 at 35–60% saturation. This hydroxylase was specific for GA1 and did not hydroxylate either pseudo-GA1-[3H] or 16-ketoGA1-[3H]. Virtually all hydroxylase activity was localized in the cotyledons.  相似文献   

18.
The metabolic fate of 1-β-d -arabinofuranosyl-5-[(E)-2-bromovinyl]uracil (BV-araU) in herpes simplex virus type 1-infected cells was studied using tritium-labeled BV-araU. [3H]BV-araU was selectively taken-up by infected cells. Approximately 10% of the total uptake of [3H]BV-araU was recovered from the acid-insoluble fraction at any time post-infection. Both cellular uptake of [3H]BV-araU and its incorporation into the acid-insoluble fraction increased with increasing incubation time through 8 hr post-infection. Uptake of [3H]BV-araU and its incorporation into the acid-insoluble fraction also increased proportionally to the duration of exposure to [3H]BV-araU. An alkaline sucrose gradient sedimentation analysis revealed that the radioactive DNA obtained from cells pulse-labeled with [3H]BV-araU were small DNA fragments which remained at the top following a chasing period in isotope-free medium, whereas that pulse-labeled with [3H]thymidine was chased to a fraction of high molecular weight DNA. Nuclease P1 digestion reduced 99% of the [3H]BV-araU-labeled DNA extracted from infected cells to a low molecular weight. Following digestion of [3H]BV-araU-labeled DNA with micrococcal nuclease and spleen exonuclease, all of the radioactivity was recovered as [3H]BV-araU 3′-monophosphate. Thus, BV-araU strongly inhibits the elongation of viral DNA strands as demonstrated by the alkaline sucrose gradient sedimentation analysis, whereas at least a portion of the [3H]BV-araU is incorporated inside viral DNA strands in infected cells.  相似文献   

19.
A quantitative assay employing binding of [3H]diisopropylfluorophosphate ([3H]DFP) and SDS-polyacrylamide gel electrophoresis was used to measure serine hydrolases in cell-free extracts from rat splenic lymphocytes. After labeling with [3H]DFP at pH 7, six major serine hydrolases are detected on 10% gels, having molecular weights of 78, 55, 34, 30, 28 and 17 (· 10?3). When labeled at pH 4, only four activities are measured, with Mr or 79, 55, 33 and 17 (· 10?3). Incubation of splenic lymphocytes for 8 h in vitro with 1 μM dexamethasone followed by [3H]DFP labeling at pH 7 produces a 91% increase in the 17000 [3H]DFP. Hormone treatment for 8 h with subsequent labeling at pH 4 results in a 15% increase in the largest (78000) species, as well as 73% increase in the 17000 enzyme, compared with lysates from cells incubated without steroid. These effects are not observed after only 4 h of glucocorticoid exposure. Dexamethasone treatment for 8 h does not produce a decrease in any of these serine hydrolases, nor is there an apparent induction of new enzymes (i.e., having a molecular weight different from the preexisting species). Studies examining the effect of protease inhibitors on the [3H]DFP capacity of these proteins, show that the 17000 enzyme is sensitive to the protease inhibitor, pepstatin A, as well as the sulfhydryl reagents dithiothreitol and N-ethylmaleimide. These result suggest that this dexamethasone-responsive enzyme is a protease which requires a free thiol group for optimal activity. These findings are discussed with regard to the mechanism of glucocorticoid action in lymphocytes.  相似文献   

20.
Nucleotides such as GTP and GDP appear to be involved in signal transduction via G protein modulation of adenylate cyclase activity. Studies on direct binding of [3H]GDP to membranes prepared from cultured immature rat Sertoli cells indicated that this process was reversible, approached steady state within 10 min, had a Ka of 4.5 ·106M−1 and was specific for guanine nucleotides. The non-hydrolyzable analog, guanosine 5′-O-[3-thio]triphosphate (GPPP[S]), was most effective as an inhibitor of [3H]GDP binding (ED50 = 4.8·10−8M), whereas guanosine 5′-O-[2-thio]diphosphate (Gpp[S]) was less potent (ED50 = 3.4·10−7M). Release of bound GDP was enhanced by follitropin (FSH) in the presence of Gppp[S], although not by FSH alone. Sertoli cell membranes possess guanine nucleotide hydrolase activity, where 95% of added nucleotide was rapidly degraded to guanosine. Binding kinetics were significantly influenced by nucleotide metabolism, which was prevented by controlling the Mg2+ concentration with EDTA and including App[NH]p to reduce nonspecific hydrolysis. Kinetic studies indicated that Gpp[S] inhibited (P < 0.05) Gppp[S]-stimulated adenylate cyclase activity (Ki = 1.8·10−7M), whereas basal activity remained unaffected. Addition of Gpp[S] to pre-activated enzyme (FSH plus GTP) resulted in a time-dependent decay of adenylate cyclase activity with a Koff value of 6 ± 1·min−1. Using a two-stage pre-inculbation technique, adenylate cyclase activity was demonstrated to be sensitive to the nucleotide bound. When FSH was included, catalytic activity was not altered by the order of pre-incubation with the nucleotides. This suggested that the exchange of bound Gpp[S] for Gppp[S] was enhance by FSH. Activation and attenuation of FSH-sensitive adenylate cyclase activity is dependent on a nucleotide exchange mechanism which is driven by (1) the higher affinity of G for GTP than GDP, (2) enhanced release of GD when FSH is present and (3) GTP hydrolysis coupled to rapid metabolism of guanine nucleotides.  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