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 468 毫秒
1.
1. all-trans-Retinoic acid at concentrations greater than 10−7m stimulated the incorporation of d-[3H]glucosamine into 8m-urea/5% (w/v) sodium dodecyl sulphate extracts of 1m-CaCl2-separated epidermis from pig ear skin slices cultured for 18h. The incorporation of 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected. 2. Electrophoresis of the solubilized epidermis showed increased incorporation of d-[3H]glucosamine into a high-molecular-weight glycosaminoglycan-containing peak when skin slices were cultured in the presence of 10−5m-all-trans-retinoic acid. The labelling of other epidermal components with d-[3H]glucosamine, 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected by 10−5m-all-trans-retinoic acid. 3. Trypsinization dispersed the epidermal cells and released 75–85% of the total d-[3H]glucosamine-labelled material in the glycosaminoglycan peak. Thus most of this material was extracellular in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 4. Increased labelling of extracellular epidermal glycosaminoglycans was also observed when human skin slices were treated with all-trans-retinoic acid, indicating a similar mechanism in both tissues. Increased labelling was also found when the epidermis was cultured in the absence of the dermis, suggesting a direct effect of all-trans-retinoic acid on the epidermis. 5. Increased incorporation of d-[3H]-glucosamine into extracellular epidermal glycosaminoglycans in 10−5m-all-trans-retinoic acid-treated skin slices was apparent after 4–8h in culture and continued up to 48h. all-trans-Retinoic acid (10−5m) did not affect the rate of degradation of this material in cultures `chased' with 5mm-unlabelled glucosamine after 4 or 18h. 6. Cellulose acetate electrophoresis at pH7.2 revealed that hyaluronic acid was the major labelled glycosaminoglycan (80–90%) in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 7. The labelling of epidermal plasma membranes isolated from d-[3H]glucosamine-labelled skin slices by sucrose density gradient centrifugation was similar in control and 10−5m-all-trans-retinoic acid-treated tissue. 8. The results indicate that increased synthesis of mainly extracellular glycosaminoglycans (largely hyaluronic acid) may be the first response of the epidermis to excess all-trans-retinoic acid.  相似文献   

2.
1. When pig ear skin slices were cultured for 18h in the presence of 1μg of tunicamycin/ml the incorporation of d-[3H]glucosamine into the epidermis, solubilized with 8m-urea/5% (w/v) sodium dodecyl sulphate, was inhibited by 45–55%. This degree of inhibition was not increased by using up to 5μg of tunicamycin/ml or by treating the skin slices with tunicamycin for up to 8 days. The incorporation of (U-14C)-labelled l-amino acids under these conditions was not affected by tunicamycin. Polyacrylamide-gel electrophoresis indicated that the labelling of the major glycosaminoglycan peak with d-[3H]glucosamine was unaffected, whereas that of the faster migrating glycoprotein components was considerably decreased in the presence of tunicamycin. 2. Subcellular fractionation indicated that tunicamycin specifically inhibited the incorporation of d-[3H]glucosamine but not of (U-14C)-labelled l-amino acids into particulate (mainly plasma-membrane) glycoproteins by about 70%. The labelling of soluble glycoproteins was hardly affected. Polyacrylamide-gel electrophoresis of the plasma-membrane fraction showed decreased d-[3H]glucosamine incorporation into all glycoprotein components, indicating that the plasma-membrane glycoproteins contained mainly N-asparagine-linked oligosaccharides. 3. Cellulose acetate electrophoresis of both cellular and extracellular glycosaminoglycans showed that tunicamycin had no significant effect on the synthesis of the major component, hyaluronic acid. However, the incorporation of both d-[3H]glucosamine and 35SO42− into sulphated glycosaminoglycans was inhibited by about 50%. This inhibition was partially overcome, at least in the cellular fraction, by 2mm-p-nitrophenyl β-d-xyloside indicating that tunicamycin-treated epidermis retained the ability to synthesize sulphated glycosaminoglycan chains. Tunicamycin may affect the synthesis and/or degradation of proteoglycan core proteins or the xylosyltransferase. 4. Electron-microscopic examination of epidermis treated with tunicamycin for up to 4 days revealed no significant changes in cell-surface morphology or in epidermal-cell adhesion. Either N-asparagine-linked carbohydrates play little role in epidermal-cell adhesion or more probably there is little turnover of these components in epidermal adhesive structures such as desmosomes and hemidesmosomes during organ culture.  相似文献   

3.
1. When NaBr, a chaotropic salt, is added, in concentrations ranging from 0.5m to 2m, to low-salt mammary cytosol, (i) age-dependent aggregation of oestrogen receptor is inhibited, (ii) the receptor sediments as a sharp peak at 4.2S on sucrose-gradient centrifugation, with complete disappearance of heavier forms, and (iii) on gel filtration with Sephadex G-200, the receptor is included in the gel matrix. On a calibrated column, the receptor has a Stokes radius of 3.7nm (±6%). 2. Because NaBr inhibits interaction of receptor with other components of cytosol, the values of the sedimentation coefficient, measured by sucrose-gradient sedimentation, and of the Stokes radius, measured by gel filtration, can be accepted with confidence. From these values, it can be computed that the oestrogen-receptor form in NaBr has a mol.wt. of 64000, with a frictional ratio of 1.4. 3. Also, inhibition of aggregation by NaBr allows a 30–90-fold purification of oestrogen receptor. Analysis of this partially purified receptor by sucrose-gradient sedimentation and gel filtration in NaBr gives the same results as for receptor in crude cytosol. On electrofocusing on a pH5–8 gradient, the partially purified oestrogen receptor focuses at pH6.2. On removal of NaBr, receptor aggregates even in this partially purified state. It seems likely that at the protein and ionic concentrations of cytoplasm in vivo, the 64000-mol.wt. receptor form is part of higher states of self- and/or hetero-association with other cytoplasmic components. 4. NaBr up to a concentration of 2m does not inhibit binding of oestrogen by receptor, nor does it decrease the affinity of the interaction (KD8.9×10−10m). The total number of binding sites in cytosol, however, decreases by approx. 10%, but this decrease may actually be the result of elimination of lower-affinity binding by non-receptor components of cytosol. 5. NaSCN, another chaotropic salt, was also tested but gave less satisfactory results with the mammary cytosol than with uterine cytosol. EDTA was omitted from the buffers because it favours aggregation of mammary oestrogen receptor. KCl (0.4m), sucrose (15%) and ZnSO4 (3mm) did not prevent aggregation of receptor.  相似文献   

4.
1. A competitive binding assay was adapted for determination of the specific binding of glucocorticoids to cytoplasmic receptors from rat thymus cells. The steroid–receptor complexes prepared by incubation of a cytoplasmic fraction from rat thymus cells with [1,2-3H2]cortisol or with [1,2,4-3H3]triamcinolone acetonide had rates of dissociation at 37°C similar to those from intact cells. 2. The cytoplasmic receptor was unstable at 3°C, but the rate of inactivation was decreased in the presence of 2.5mm-EDTA. The steroid–receptor complex was stable. 3. Rate constants for association and for dissociation, and association constants, were determined for the interactions of cortisol, cortexolone, dexamethasone and triamcinolone acetonide with the cytoplasmic receptor at 3°C. Differences in the association constants for different steroids could largely be accounted for by the differences in the rate constants for dissociation, but the rate constants for association did not vary greatly; the implications of these findings for the nature of the steroid-binding site are discussed. 4. A cytoplasmic fraction prepared from cells which had been incubated at 37°C under anaerobic conditions bound much less [1,2-3H2]cortisol than did a fraction from aerobic cells, but the binding capacity was restored after exposure of the anaerobic cells to O2. 5. The specific binding of [1,2-3H2]-cortisol to intact thymus cells incubated aerobically was not affected by the presence of 0.1mm-cycloheximide, nor did this concentration of cycloheximide inhibit the recovery of specific binding observed when anaerobic cells were transferred to an aerobic atmosphere. 6. The energy dependence of specific binding of cortisol to the receptor is discussed with reference to possible mechanisms.  相似文献   

5.
1. Studies on the incorporation of [3-14C]pyruvate and d-3-hydroxy[3-14C]butyrate into the brain lipid fraction by brain homogenates of the suckling (7-day-old) rat have been carried out. 2. Whereas approximately twice as much CO2 was evolved from pyruvate compared with 3-hydroxybutyrate metabolism, similar amounts of the radioactivity of these two precursors were incorporated into the lipid fraction. Furthermore, in both cases the incorporation into lipid was almost tripled when glucose (10mm) or NADPH (2.5mm) was added to the incubation media. 3. If 5mm-(—)-hydroxycitrate, an ATP–citrate lyase inhibitor, was added to the incubation the incorporation of carbon from pyruvate was inhibited to 39% of the control and from 3-hydroxybutyrate to 73% of the control, whereas CO2 production from both precursors was not affected. 4. The incorporation from pyruvate or 3-hydroxybutyrate into lipids was not affected by the presence of 10mm-glutamate in the medium (to encourage N-acetylaspartate production). However, incorporation from pyruvate was inhibited by 21% in the presence of 5mm-amino-oxyacetate (a transaminase inhibitor) and by 83% in the presence of both hydroxycitrate (5mm) and amino-oxyacetate. 5. Incorporation from 3-hydroxybutyrate into brain lipids was inhibited by 20% by amino-oxyacetate alone, but by 55% in the presence of both hydroxycitrate and amino-oxyacetate. 6. It is concluded that the mechanism of carbon transfer from pyruvate into lipids across the mitochondrial membrane in the suckling rat brain is mainly via citrate and N-acetylaspartate. 3-Hydroxybutyrate, in addition to using these routes, may also be incorporated via acetoacetate formation and transport to the cytosol.  相似文献   

6.
1. The aerobic transport of d-glucose and d-galactose in rabbit kidney tissue at 25° was studied. 2. In slices forming glucose from added substrates an accumulation of glucose against its concentration gradient was found. The apparent ratio of intracellular ([S]i) and extracellular ([S]o) glucose concentrations was increased by 0·4mm-phlorrhizin and 0·3mm-ouabain. 3. Slices and isolated renal tubules actively accumulated glucose from the saline; the apparent [S]i/[S]o fell below 1·0 only at [S]o higher than 0·5mm. 4. The rate of glucose oxidation by slices was characterized by the following parameters: Km 1·16mm; Vmax. 4·5μmoles/g. wet wt./hr. 5. The active accumulation of glucose from the saline was decreased by 0·1mm-2,4-dinitrophenol, 0·4mm-phlorrhizin and by the absence of external Na+. 6. The kinetic parameters of galactose entry into the cells were: Km 1·5mm; Vmax 10μmoles/g. wet wt./hr. 7. The efflux kinetics from slices indicated two intracellular compartments for d-galactose. The galactose efflux was greatly diminished at 0°, was inhibited by 0·4mm-phlorrhizin, but was insensitive to ouabain. 8. The following mechanism of glucose and galactose transport in renal tubular cells is suggested: (a) at the tubular membrane, these sugars are actively transported into the cells by a metabolically- and Na+-dependent phlorrhizin-sensitive mechanism; (b) at the basal cell membrane, these sugars are transported in accordance with their concentration gradient by a phlorrhizin-sensitive Na+-independent facilitated diffusion. The steady-state intracellular sugar concentration is determined by the kinetic parameters of active entry, passive outflow and intracellular utilization.  相似文献   

7.
The effects of aliphatic 2-oxocarboxylic acids, at concentrations of up to 40mm, on the function of pancreatic islets from ob/ob (obese–hyperglycaemic) mice were investigated. 1. 2-Oxopentanoate, dl-3-methyl-2-oxopentanoate, 4-methyl-2-oxopentanoate and 2-oxohexanoate all induced insulin release by isolated incubated islets and a biphasic insulin-secretory pattern in perfused mouse pancreas. The last two substances were similar in potency to glucose. Pyruvate, 2-oxobutyrate, 3-methyl-2-oxobutyrate and 2-oxo-octanoate did not induce insulin release significantly. 2. 2-Oxocarboxylic acids with significant insulin-secretory potency also induced significant 45Ca uptake by isolated incubated islets. 3. The rates of decarboxylation of [1-14C]pyruvate, 3-methyl-2-oxo[1-14C]butyrate and 4-methyl-2-oxo[1-14C]pentanoate were twice as high as the rates of oxidation of the corresponding U-14C-labelled compounds. However, whereas the rates of metabolism of labelled pyruvate and 3-methyl-2-oxobutyrate steadily increased over the concentration range 1–40mm, those of labelled 4-methyl-2-oxopentanoate and d-[U-14C]glucose levelled off at concentrations above 10mm. 4. Omission of 40CaCl2 from the incubation medium reduced the rate of oxidation of the insulin secretagogue [U-14C]4-methyl-2-oxopentanoate, but left that of the non-(insulin secretagogue) [U-14C]3-methyl-2-oxobutyrate unaffected. 5. Only glucose, and not pyruvate, 3-methyl-2-oxobutyrate and 4-methyl-2-oxopentanoate, significantly inhibited oxidation of endogenous fatty acids. 6. It is suggested that stimulus–secretion coupling and the resulting exocytosis of insulin in pancreatic β-cells may modulate both fuel oxidation and 45Ca uptake.  相似文献   

8.
1. The hormonal control of glycogen breakdown was studied in hepatocytes isolated from livers of fed rats. 2. Glucose release was stimulated by [8-arginine]vasopressin (10pm–10nm), oxytocin (1nm–1μm), and angiotensin II (1nm–0.1μm). These responses are all at least as sensitive to hormone as is glucose output in the perfused rat liver. 3. The effect of these three hormones on glucose release was critically dependent on extracellular Ca2+, unlike that of glucagon. Half-maximal restoration of the vasopressin response occurred if 0.3mm-Ca2+ was added back to the incubation medium. 4. Glycogen breakdown was more than sufficient to account for the glucose released into the medium, in the absence or presence of hormones. Lactate release by hepatocytes was not affected by vasopressin, but was inhibited by glucagon. 5. If Ca2+ was omitted from the extracellular medium, vasopressin stimulated glycogenolysis, but not glucose release. 6. The phosphorylase a content of hepatocytes was increased by vasopressin, oxytocin and angiotensin II; minimum effective concentrations were 0.1pm, 0.1nm and 10pm respectively. This response was also dependent on Ca2+. 7. These results demonstrate that hepatocytes can respond to low concentrations of vasopressin and angiotensin II, i.e. these effects are likely to be relevant in the intact animal. The role of extracellular Ca2+ in the effects of these hormones on hepatic glycogenolysis and glucose release is discussed.  相似文献   

9.
1. Phosphomevalonate kinase and 5-pyrophosphomevalonate decarboxylase have been purified from the freeze-dried latex serum of the commercial rubber tree Hevea brasiliensis. 2. The phosphomevalonate kinase was acid- and heat-labile and required the presence of a thiol to maintain activity. 3. The 5-pyrophosphomevalonate decarboxylase was relatively acid-stable and more heat-stable than the phosphokinase. 4. Maximum activity of the phosphokinase was achieved at pH 7.2 with 0.2mm-5-phosphomevalonate (Km 0.042mm), 2.0mm-ATP (Km 0.19mm) and 8mm-Mg2+ at 40°C. The apparent activation energy was 14.8kcal/mol. 5. Maximum activity of 5-pyrophosphomevalonate decarboxylase was achieved at pH5.5–6.5 with 0.1mm-5-pyrophosphomevalonate (Km 0.004mm), 1.5mm-ATP (Km 0.12mm) and 2mm-Mg2+. The apparent activation energy was 13.7kcal/mol. The enzyme was somewhat sensitive to inhibition by its products, isopentenyl pyrophosphate and ADP.  相似文献   

10.
Potassium fluxes in dialyzed squid axons   总被引:11,自引:6,他引:5       下载免费PDF全文
Measurements have been made of K influx in squid giant axons under internal solute control by dialysis. With [ATP]i = 1 µM, [Na]i = 0, K influx was 6 ± 0.6 pmole/cm2 sec; an increase to [ATP]i = 4 mM gave an influx of 8 ± 0.5 pmole/cm2 sec, while [ATP]i 4, [Na]i 80 gave a K influx of 19 ± 0.7 pmole/cm2 sec (all measurements at ∼16°C). Strophanthidin (10 µM) in seawater quantitatively abolished the ATP-dependent increase in K influx. The concentration dependence of ATP-dependent K influx on [ATP]i, [Na]i, and [K]o was measured; an [ATP]i of 30 µM gave a K influx about half that at physiological concentrations (2–3 mM). About 7 mM [Na]i yielded half the K influx found at 80 mM [Na]i. The ATP-dependent K influx responded linearly to [K]o from 1–20 mM and was independent of whether Na, Li, or choline was the principal cation of seawater. Substances tested as possible energy sources for the K pump were acetyl phosphate, phosphoarginine, PEP, and d-ATP. None was effective except d-ATP and this substance gave 70% of the maximal flux only when phosphoarginine or PEP was also present.  相似文献   

11.
1. The formation of adenosine 5′-phosphate, guanosine 5′-phosphate and inosine 5′-phosphate from [8-14C]adenine, [8-14C]guanine and [8-14C]hypoxanthine respectively in the presence of 5-phosphoribosyl pyrophosphate and an extract from Ehrlich ascites-tumour cells was assayed by a method involving liquid-scintillation counting of the radioactive nucleotides on diethylaminoethylcellulose paper. The results obtained with guanine were confirmed by a spectrophotometric assay which was also used to assay the conversion of 6-mercaptopurine and 5-phosphoribosyl pyrophosphate into 6-thioinosine 5′-phosphate in the presence of 6-mercaptopurine phosphoribosyltransferase from these cells. 2. At pH 7·8 and 25° the Michaelis constants for adenine, guanine and hypoxanthine were 0·9 μm, 2·9 μm and 11·0 μm in the assay with radioactive purines; the Michaelis constant for guanine in the spectrophotometric assay was 2·6 μm. At pH 7·9 the Michaelis constant for 6-mercaptopurine was 10·9 μm. 3. 25 μm-6-Mercaptopurine did not inhibit adenine phosphoribosyltransferase. 6-Mercaptopurine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 4·7 μm) and hypoxanthine phosphoribosyltransferase (Ki 8·3 μm). Hypoxanthine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 3·4 μm). 4. Differences in kinetic parameters and in the distribution of phosphoribosyltransferase activities after electrophoresis in starch gel indicate that different enzymes are involved in the conversion of adenine, guanine and hypoxanthine into their nucleotides. 5. From the low values of Ki for 6-mercaptopurine, and from published evidence that ascites-tumour cells require supplies of purines from the host tissues, it is likely that inhibition of hypoxanthine and guanine phosphoribosyltransferases by free 6-mercaptopurine is involved in the biological activity of this drug.  相似文献   

12.
Effect of adrenaline on 32P incorporation into rat fat-cell phospholipids   总被引:3,自引:3,他引:0  
1. The phospholipid composition of fat-cells prepared from rat epididymal fat-pad was determined. 2. The incorporation of [32P]Pi into the phospholipids of fat-cells incubated in glucose-free medium and the effect of adrenaline and of α- and β-adrenergic blocking agents, were studied. 3. Incorporation of [32P]Pi into fat-cell phospholipid increased with time; incubation with adrenaline resulted in increased incorporation that was related to the concentration of adrenaline. 4. The pattern of incorporation of [32P]Pi into the individual phospholipids of fat-cells after incubation for 1h was determined; adrenaline (5.4μm) resulted in increased incorporation into phosphatidylcholine. 5. Incubation of fat-cells with propranolol (34μm) and adrenaline (5.4μm) resulted in abolition of adrenaline-stimulated lipolysis; there was a decrease in the specific radioactivity of phosphatidylcholine and an increase in the specific radioactivity of phosphatidylethanolamine, phosphatidic acid, phosphatidylinositol and cardiolipin compared with cells incubated with adrenaline alone. 6. Incubation of fat-cells with phenoxybenzamine (0.1mm) and adrenaline (5.4μm) resulted in stimulation of lipolysis, and in diminished specific radioactivities of phosphatidylcholine, phosphatidic acid, phosphatidylinositol, phosphatidylglycerol and choline plasmalogen compared with cells stimulated with adrenaline alone.  相似文献   

13.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

14.
Both cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase were recovered mainly from the supernatant fractions of guinea-pig pancreas, but a higher proportion of the activity of the former was associated with the pellet fractions. The activities in the supernatant were not separated by gel filtration, but were clearly separated by subsequent chromatography on an anion-exchange resin. The activities of cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase had high-affinity (Km 6.5±1.1μm and 31.9±3.9μm respectively) and low-affinity (Km 0.56±0.05mm and 0.32±0.03mm respectively) components. The activity of neither enzyme was affected by the pancreatic secretogens, cholecystokinin-pancreozymin, secretin and carbachol. Removal of ions by gel filtration resulted in a marked reduction in cyclic nucleotide phosphodiesterase activity, which could be restored by addition of Mg2+. Mn2+ (3mm) was as effective as Mg2+ (3mm) in the case of cyclic AMP phosphodiesterase, but was less than half as effective in the case of cyclic GMP phosphodiesterase. The metal-ion chelators, EDTA and EGTA, also decreased activity. Ca2+ (1mm) did not affect the activity of cyclic nucleotide phosphodiesterase when the concentration of Mg2+ was 3mm. At concentrations of Mg2+ between 0.1 and 1mm, 1mm-Ca2+ was activatory, and at concentrations of Mg2+ below 0.1mm, 1mm-Ca2+ was inhibitory. These results are discussed in terms of the possible significance of cyclic nucleotide phosphodiesterase in the physiological control of cyclic nucleotide concentrations during stimulus–secretion coupling.  相似文献   

15.
1. The `30s' and `50s' ribosomes from ribonuclease-active (Escherichia coli B) and -inactive (Pseudomonas fluorescens and Escherichia coli MRE600) bacteria have been studied in the ultracentrifuge. Charge anomalies were largely overcome by using sodium chloride–magnesium chloride solution, I 0·16, made 0–50mm with respect to Mg2+. 2. Differentiation of enzymic and physical breakdown at Mg2+ concentrations less than 5mm was made by comparing the properties of E. coli B and P. fluorescens ribosomes. 3. Ribonuclease-active ribosomes alone showed a transformation of `50s' into 40–43s components. This was combined with the release of a small amount of `5s' material which may be covalently bound soluble RNA. Other transformations of the `50s' into 34–37s components were observed in both ribonuclease-active and -inactive ribosomes at 1·0–2·5mm-Mg2+, and also with E. coli MRE600 when EDTA (0·2mm) was added to a solution in 0·16m-sodium chloride. 4. Degradation of ribonuclease-active E. coli B ribosomes at Mg2+ concentration 0·25mm or less was coincident with the formation of 16s and 21s ribonucleoprotein in P. fluorescens, and this suggested that complete dissociation of RNA from protein was not an essential prelude to breakdown of the RNA by the enzyme. 5. As high Cs+/Mg2+ ratios cause ribosomal degradation great care is necessary in the interpretation of equilibrium-density-gradient experiments in which high concentrations of caesium chloride or similar salts are used. 6. The importance of the RNA moiety in understanding the response of ribosomes to their ionic environment is discussed.  相似文献   

16.
Hyperinsulinemia (HI) is elevated plasma insulin at basal glucose. Impaired glucose tolerance is associated with HI, although the exact cause and effect relationship remains poorly defined. We tested the hypothesis that HI can result from an intrinsic response of the β-cell to chronic exposure to excess nutrients, involving a shift in the concentration dependence of glucose-stimulated insulin secretion. INS-1 (832/13) cells were cultured in either a physiological (4 mm) or high (11 mm) glucose concentration with or without concomitant exposure to oleate. Isolated rat islets were also cultured with or without oleate. A clear hypersensitivity to submaximal glucose concentrations was evident in INS-1 cells cultured in excess nutrients such that the 25% of maximal (S0.25) glucose-stimulated insulin secretion was significantly reduced in cells cultured in 11 mm glucose (S0.25 = 3.5 mm) and 4 mm glucose with oleate (S0.25 = 4.5 mm) compared with 4 mm glucose alone (S0.25 = 5.7 mm). The magnitude of the left shift was linearly correlated with intracellular lipid stores in INS-1 cells (r2 = 0.97). We observed no significant differences in the dose responses for glucose stimulation of respiration, NAD(P)H autofluorescence, or Ca2+ responses between left- and right-shifted β-cells. However, a left shift in the sensitivity of exocytosis to Ca2+ was documented in permeabilized INS-1 cells cultured in 11 versus 4 mm glucose (S0.25 = 1.1 and 1.7 μm, respectively). Our results suggest that the sensitivity of exocytosis to triggering is modulated by a lipid component, the levels of which are influenced by the culture nutrient environment.  相似文献   

17.
1. Rat tissue homogenates convert dl-1-aminopropan-2-ol into aminoacetone. Liver homogenates have relatively high aminopropanol-dehydrogenase activity compared with kidney, heart, spleen and muscle preparations. 2. Maximum activity of liver homogenates is exhibited at pH9·8. The Km for aminopropanol is approx. 15mm, calculated for a single enantiomorph, and the maximum activity is approx. 9mμmoles of aminoacetone formed/mg. wet wt. of liver/hr.at 37°. Aminoacetone is also formed from l-threonine, but less rapidly. An unidentified amino ketone is formed from dl-4-amino-3-hydroxybutyrate, the Km for which is approx. 200mm at pH9·8. 3. Aminopropanol-dehydrogenase activity in homogenates is inhibited non-competitively by dl-3-hydroxybutyrate, the Ki being approx. 200mm. EDTA and other chelating agents are weakly inhibitory, and whereas potassium chloride activates slightly at low concentrations, inhibition occurs at 50–100mm. 4. It is concluded that aminopropanol-dehydrogenase is located in mitochondria, and in contrast with l-threonine dehydrogenase can be readily solubilized from mitochondrial preparations by ultrasonic treatment. 5. Soluble extracts of disintegrated mitochondria exhibit maximum aminopropanol-dehydrogenase activity at pH9·1 At this pH, Km values for the amino alcohol and NAD+ are approx. 200 and 1·3mm respectively. Under optimum conditions the maximum velocity is approx. 70mμmoles of aminoacetone formed/mg. of protein/hr. at 37°. Chelating agents and thiol reagents appear to have little effect on enzyme activity, but potassium chloride inhibits at all concentrations tested up to 80mm. dl-3-Hydroxybutyrate is only slightly inhibitory. 6. Dehydrogenase activities for l-threonine and dl-4-amino-3-hydroxybutyrate appear to be distinct from that for aminopropanol. 7. Intraperitoneal injection of aminopropanol into rats leads to excretion of aminoacetone in the urine. Aminoacetone excretion proportional to the amount of the amino alcohol administered, is complete within 24hr., but represents less than 0·1% of the dose given. 8. The possible metabolic role of amino alcohol dehydrogenases is discussed.  相似文献   

18.
Citrulline formation by both human neuronal nitric-oxide synthase (nNOS) and mouse macrophage inducible NOS was inhibited by the hydrogen sulfide (H2S) donor Na2S with IC50 values of ∼2.4·10−5 and ∼7.9·10−5 m, respectively, whereas human endothelial NOS was hardly affected at all. Inhibition of nNOS was not affected by the concentrations of l-arginine (Arg), NADPH, FAD, FMN, tetrahydrobiopterin (BH4), and calmodulin, indicating that H2S does not interfere with substrate or cofactor binding. The IC50 decreased to ∼1.5·10−5 m at pH 6.0 and increased to ∼8.3·10−5 m at pH 8.0. Preincubation of concentrated nNOS with H2S under turnover conditions decreased activity after dilution by ∼70%, suggesting irreversible inhibition. However, when calmodulin was omitted during preincubation, activity was not affected, suggesting that irreversible inhibition requires both H2S and NO. Likewise, NADPH oxidation was inhibited with an IC50 of ∼1.9·10−5 m in the presence of Arg and BH4 but exhibited much higher IC50 values (∼1.0–6.1·10−4 m) when Arg and/or BH4 was omitted. Moreover, the relatively weak inhibition of nNOS by Na2S in the absence of Arg and/or BH4 was markedly potentiated by the NO donor 1-(hydroxy-NNO-azoxy)-l-proline, disodium salt (IC50 ∼ 1.3–2.0·10−5 m). These results suggest that nNOS and inducible NOS but not endothelial NOS are irreversibly inhibited by H2S/NO at modest concentrations of H2S in a reaction that may allow feedback inhibition of NO production under conditions of excessive NO/H2S formation.  相似文献   

19.
Functional expression in heterologous hosts is often less successful for integral membrane proteins than for soluble proteins. Here, two Ambrosiozyma monospora transporters were successfully expressed in Saccharomyces cerevisiae as tagged proteins. Growth of A. monospora on l-arabinose instead of glucose caused transport activities of l-arabinose, l-arabitol, and ribitol, measured using l-[1-3H]arabinose, l-[14C]arabitol, and [14C]ribitol of demonstrated purity. A. monospora LAT1 and LAT2 genes were cloned earlier by using their ability to improve the growth of genetically engineered Saccharomyces cerevisiae on l-arabinose. However, the l-arabinose and pentitol transport activities of S. cerevisiae carrying LAT1 or LAT2 are only slightly greater than those of control strains. S. cerevisiae carrying the LAT1 or LAT2 gene fused in frame to the genes for green fluorescent protein (GFP) or red fluorescent protein (mCherry) or adenylate kinase (AK) exhibited large (>3-fold for LAT1; >20-fold for LAT2) increases in transport activities. Lat1-mCherry transported l-arabinose with high affinity (Km ≈ 0.03 mM) and l-arabitol and ribitol with very low affinity (Km ≥ 75 mM). The Lat2-GFP, Lat2-mCherry, and Lat2-AK fusion proteins could not transport l-arabinose but were high-affinity pentitol transporters (Kms ≈ 0.2 mM). The l-arabinose and pentitol transport activities of A. monospora could not be completely explained by any combination of the observed properties of tagged Lat1 and Lat2, suggesting either that tagging and expression in a foreign membrane alters the transport kinetics of Lat1 and/or Lat2 or that A. monospora contains at least one more l-arabinose transporter.  相似文献   

20.
A protein-binding radioassay for cyclic AMP was modified to detect less than 0.025pmol of the nucleotide. The method was applied to the measurement of cyclic AMP in small numbers of mouse pancreatic islets (as little as 25μg of tissue) by use of barium acetate–H2SO4 for deproteinization. The concentration of cyclic AMP in mouse islets incubated in media containing 3.3 or 20mm-glucose was 0.016pmol/10 islets (approx. 1μm in intracellular water). Glucose concentration (3.3 or 20mm) had no detectable effect on islet concentrations of cyclic AMP with periods of incubation or perifusion ranging from 0.5 to 60min, although insulin release rate was rapidly increased by 20mm-glucose. Caffeine (5mm) or 3-isobutyl-1-methylxanthine (1mm), which are known inhibitors of islet cyclic AMP phosphodiesterase, produced marked and rapid increases in islet cyclic AMP concentration at 3.3 or 20mm-glucose, but only enhanced the insulin release rate at the higher glucose concentration. The role of cyclic AMP in insulin release induced by glucose 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