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.
Summary The activity of acid phosphatase in skeletal muscle fibres of the plantaris and soleus of normal and dystrophic male hamsters was quantified using a histochemical post-coupling semipermeable membrane technique. Althoug the absolute levels of activity were found to vary widely from one animal to another, the ratio of the mean activities in the two muscles in each animal was virtually constant. In normal muscles, the ratio was about 0.73 and in dystrophic muscles, about 0.77. The activity in plantaris muscle fibres was always significantly lower than that in the corresponding soleus fibres, and in normal fibres compared to dystrophic ones. Another difference was that in normal fibres the mean activity declined to a constant level in mature animals older than about 3 months. In contrast, the activity in dystrophic muscles appeared to fall exponentially throughout life. The functional significance of these findings is discussed.In honour of Prof. P. van Duijn  相似文献   

2.
The activity of acid phosphatase in skeletal muscle fibres of the plantaris and soleus of normal and dystrophic male hamsters was quantified using a histochemical post-coupling semipermeable membrane technique. Although the absolute levels of activity were found to vary widely from one animal to another, the ratio of the mean activities in the two muscles in each animal was virtually constant. In normal muscles, the ratio was about 0.73 and in dystrophic muscles, about 0.77. The activity in plantaris muscle fibres was always significantly lower than that in the corresponding soleus fibres, and in normal fibres compared to dystrophic ones. Another difference was that in normal fibres the mean activity declined to a constant level in mature animals older than about 3 months. In contrast, the activity in dystrophic muscles appeared to fall exponentially throughout life. The functional significance of these findings is discussed.  相似文献   

3.
Adenyl cyclase and cyclic nucleotide phosphodiesterase activities were assayed in homogenates of hind leg skeletal muscle from dystrophic and normal mice. Adenyl cyclase activity was stimulated 2.5 times by epinephrine and 6 times by fluoride over the basal activity in both dystrophic and normal mice. The activity of adenyl cyclase from dystrophic muscle of mice was significantly higher than that of normal mice under all the conditions tested (i.e. basal, epinephrine and fluoride). Cyclic nucleotide phosphodiesterase from skeletal muscle of mice has two Km's (2.1 and 11 mumol/l) which suggests the existence of either two forms of enzyme or a single enzyme with negative cooperativity. The activity of this enzyme was significantly elevated in the skeletal muscle of dystrophic mice compared to the normal controls. The available evidence suggests that the same cyclic nucleotide phosphodiesterase is responsible for the hydrolysis of both cyclic AMP and cyclic GMP.  相似文献   

4.
Polyamine (putrescine, spermidine, and spermine) contents and ornithine (ODC) and S-adenosylmethionine (SAMDC) decarboxylase activities have been assessed in an age-dependent manner, in normal and polymyopathic (dystrophic) hamster skeletal muscle, heart, and tongue extract and in primary tongue myoblast and skin fibroblast cultures. At 2 weeks of age, polyamine contents were significantly elevated in all of the dystrophic hamster tissues studied when compared with their age-matched controls. The degree of this elevation decreased with the age of the animals, generally, to a level where no significant difference in polyamine contents could be noted between normal and dystrophic hamster tissues. ODC and SAMDC activities in whole tissue extracts were consistently highest in 2-week-old muscle extracts and also declined with age. However, no significant changes in ODC or SAMDC activities were evident in any of the dystrophic muscle tissues studied when compared with their age-matched controls. Polyamine contents in dystrophic hamster myoblast and fibroblast primary cultures were also during proliferation (1 and 2 days after the initial seeding) compared with cultures prepared from normal hamsters. ODC and SAMDC activities in primary myoblast and fibroblast cultures clearly reflected the rate of cell proliferation, with highest activities found in subconfluent cell cultures. However, in general, no significant dystrophic-related abnormality in ODC or SAMDC activity was evident in proliferating myoblast or fibroblast cultures. These results suggest that the elevated polyamine contents of dystrophic hamster tissues and primary cultures may be due to a deficiency in polyamine catabolism or transport.  相似文献   

5.
Polysomes from the skeletal muscle of normal and dystrophic hamsters were dissociated into ribosomal subunits by treatment with puromycin and the subunits from both strains were reassociated in all possible combinations. When their protein synthesis activity was assayed in a poly(U)-directed cell-free system at a low magnesium concentration, the reassociated ribosomes from dystrophic hamsters were less active than the ribosomes from control animals. The ribosomal defect is a property of the 60S subunit and is due to a ribosomal component rather than to abnormal binding of a non-ribosomal protein.  相似文献   

6.
The rates of loss of adenylate kinase and creatine kinase from the circulation after intravenous injection of homogenous chicken skeletal muscle enzymes were examined to determine the role of plasma clearance rates in determining the plasma levels of these enzymes in normal and dystrophic chickens. The rapid clearance of adenylate kinase activity (average half-life of 5 min) and the slower biphasic clearance of creatine kinase activity (average half-lives of 0.95 and 11 hr) are consistent with the elevation of creatine kinase but not adenylate kinase in the blood plasma of dystrophic chickens compared to normal chickens. The rates of clearance of these enzymes were similar in normal chickens compared to dystrophic chickens. Radioiodinated enzymes were cleared at similar, but slightly more rapid rates than the loss of enzyme activity. The loss of adenylate kinase activity from the circulation may be due in part to inactivation since adenylate kinase activity is rapidly inactivated in serum in vitro, and because no increase in adenylate kinase activity is observed in the most specific sites of clearance of the radioiodinated enzyme, the liver and spleen. The comparison of enzyme activities in press juices to the activities in high-ionic-strength homogenates of muscle tissue from normal and dystrophic muscle, indicates that adenylate kinase activity is not associated with intracellular structures to the extent that would prohibit release from dystrophic muscle tissue. These results, and those presented previously with regard to plasma levels and clearance rates of AMP aminohydrolase and pyruvate kinase in normal and dystrophic chickens (11) support our hypothesis that the rates of loss of muscle enzyme activities from the circulation are important in determining the circulating levels of muscle enzymes in dystrophic chickens. Furthermore, from the measurement of plasma levels and clearance rates of creatine kinase, it was estimated that the efflux rate of creatine kinase from dystrophic muscle tissue is 2.0% of the total breast muscle creatine kinase per day.  相似文献   

7.
Isozyme patterns of cholinesterase (ChE) from heart, tongue, and skeletal muscle of normal and dystrophic hamsters are presented. Two principal bands, bands 1 and 2, were evaluated. Band 1 migrates faster towards the anode than does band 2. While bands 1 and 2 stain for AChE and were found in control muscles, only band 2 was stained by a pseudocholinesterase (BuChE) and was decreased in samples from dystrophic hamsters. The decrease in BuChE was most pronounced in dystrophic heart muscle. The low level of BuChE measured for dystrophic animal tissue was similar to isozyme patterns found in embryonic tissue and in denervated muscle. BuChE obtained by acrylamide gel electrophoresis along with 16S AchE appears to be a useful biochemical marker of nerve-muscle interactions.  相似文献   

8.
Cyclic AMP content, adenylate cyclase (EC 4.6.1.1) activity and phosphodiesterase I (EC 3.1.4.1) activity of the hind leg skeletal muscle and cardiac muscle in 60- and 150-day-old normal and myopathic (UM-X7.1) hamsters were examined. In 60-day-old myopathic animals, cardiac cyclic AMP levels were higher and phosphodiesterase I activity was lower, without any changes in the basal adenylate cyclase activity, whereas in 150-day-old myopathic hamsters, cardiac cyclic AMP and basal adenylate cyclase activity were lower, without any changes in the homogenate phosphodiesterase I activity. On the other hand, basal adenylate cyclase and phosphodiesterase I activities in the skeletal muscle homogenate from 60- and 150-day-old myopathic animals were not different from the normal values but the skeletal muscle cyclic AMP levels were significantly less in 60-day-old myopathic hamsters only. The plasma cyclic AMP levels in 60-day-old myopathic hamsters, unlike 150-day-old myopathic animals, were higher than the normal. Although these results reveal differences in myopathic cardiac and skeletal muscles, it is concluded that changes in adenylate cyclase-cyclic AMP system in myopathy are dependent upon the degree of disease.  相似文献   

9.
The activity and diurnal variation of 3-hydroxy-3-methyglutaryl-CoA reductase (EC 1.1.1.34; HMG-CoA reductase), the rate-limiting enzyme in the cholesterol-biosynthetic pathway, of normal and dystrophic hamsters was determined. Liver enzyme activity showed a diurnal pattern in the normal male, but not in the dystrophic male. Enzyme values in normal males at the midpoint of the 12 h dark period were 10 times those in dystrophic males. No evidence for diurnal variation in the HMG-CoA reductase of the brain was observed, and similar activities were found for normal and dystrophic animals. The apparent Km for HMG-CoA reductase from the liver of normal or dystrophic hamsters was approx. 9 microM, and the Vmax. was 5.9 and 21.7 pmol/min per mg of protein for dystrophic and normal hamsters respectively.  相似文献   

10.
The effects of hypothyroidism on glycogen metabolism in rat skeletal muscle were studied using the perfused rat hindlimb preparation. Three weeks after propylthiouracil treatment, serum thyroxine was undetectable and muscle glycogen and Glc-6-P were decreased. Basal and epinephrine-stimulated phosphorylase a and phosphorylase b kinase activities were also significantly reduced, as were epinephrine-stimulated cAMP accumulation and cAMP-dependent protein kinase activity. Conversely, basal and epinephrine-stimulated glycogen synthase I activities were significantly higher while the Ka of the enzyme for Glc-6-P was lower in hypothyroid animals. Propylthiouracil-treated rats also had increased phosphoprotein phosphatase activities towards phosphorylase and glycogen synthase and decreased activity of phosphatase inhibitor 1. beta-Adrenergic receptor binding and basal and epinephrine-stimulated adenylate cyclase activities were reduced in muscle particulate fractions from hypothyroid rats. Administration of triiodothyronine to rats for 3 days after 3 weeks of propylthiouracil treatment restored the altered metabolic parameters to normal. It is proposed that the decreased beta-adrenergic responsiveness of the enzymes of glycogen metabolism in hypothyroid rat skeletal muscle is due to increased activity of phosphoprotein phosphatases and to reduced beta-adrenergic receptors and adenylate cyclase activity.  相似文献   

11.
Ca2+ ATPase and calcium binding proteins were studied in cardiac and skeletal muscles of normal and dystrophic mice. In normal and dystrophic mice, Ca2+ ATPase was quite reduced in cardiac muscle compared to skeletal muscle and was, unlike skeletal muscle, insensitive to orthovanadate. Ca2+ ATPase in skeletal muscle of dystrophic mice was reduced as compared to normal mice. In both cases (normal and dystrophic), calcium binding proteins were the same (identical molecular weight). The effect of 2 drugs (Polymixine B and Bepridil) which decrease protein bound calcium was studied: the muscle proteins of dystrophic mice did not present the same sensitivity to Bepridil as controls. These findings suggest the existence of a calcium-related defect in skeletal and cardiac muscle of dystrophic mice.  相似文献   

12.
Thiol protease and cathepsin D activities were studied in extracts from hindlimb muscle of 60-day-old normal and dystrophic mice, strain 129 ReJ, and from cultured normal and dystrophic cells. Total thiol protease activity in dystrophic muscle extracts was 3.5 times higher than in normal muscle extracts, while cathepsin D, activity was 2.2 times greater in dystrophic muscle compared with normal muscle. Activation (pH 4.5, 30 degrees C) of latent thiol protease activity in extracts of muscle occurred concomitant with the inactivation or dissociation of endogenous protease inhibitors. Thiol protease assays revealed a higher ratio of active to inactive protease activity in extracts from dystrophic muscle than from normal muscle. Cultured myoblasts (L69/1) were found to contain 30-fold more thiol protease(s) and 6-fold more cathepsin D activity than whole muscle. Cells established from dystrophic muscle and grown in culture for periods up to 6 months were more responsive to thiol protease activation conditions than similar cultures derived from normal muscle. From data on the rate and extent of thiol protease activation in extracts from dystrophic cells and hindlimb muscle compared with normal tissue, it appears that cells and tissues from dystrophic mice contain a lower level of protease inhibitors than cells and tissues from normal mice.  相似文献   

13.
Previously administration of thyroxine (T4) to dystrophic hamsters improved ventilation and slowed the progression of the disease. We hypothesized that the normalization of ventilation in these animals was due to T4 improving structural and functional characteristics of the diaphragm. In the present study, contractile characteristics of the diaphragm and the extensor digitorum longus (EDL) from normal and dystrophic hamsters were evaluated after two months of T4 treatment. Compared to their placebo-treated counterparts, diaphragms and EDLs of T4-treated normal hamsters showed increased optimal muscle lengths and twitch tension, decreased contraction times and increased fatigability. T4-treatment in dystrophic hamsters showed only an increase in diaphragmatic twitch tension development. Force-frequency curves before treatment were generally higher for the normal compared to dystrophic diaphragms and EDLs. T4 administration only increased the force in normal diaphragms at the lower frequencies and in the EDLs at the higher frequencies. Although T4 serum levels were increased in both T4-treated groups, triiodothyronine (T3) was much lower in the dystrophic compared to normal hamsters, suggesting that conversion of T4 to T3 was reduced in dystrophic hamsters. We conclude that the limited functional changes in the diaphragms of T4-treated dystrophic hamsters cannot account for the marked improvement in ventilation previously reported.  相似文献   

14.
The cellular electrical activity of diaphragm from F1B normal and BIO 14.6 dystrophic hamsters has been investigated using microelectrodes. Resting membrane potentials and action potentials were recorded from control muscles and from muscles exposed to 2,4-dinitrophenol. The action potentials of normal and dystrophic diaphragms were similar in amplitude and configuration. Treatment with 2,4-dinitrophenol caused the action potential amplitude of both diaphragms to decline by similar amounts. The control resting membrane potential of diaphragm from dystrophic hamsters is not significantly different from that of normal hamsters. Treatment with 2,4-dinitrophenol caused a linear decrease in the resting membrane potentials of both groups of muscles. Dystrophic muscle, however, showed a more rapid decline in excitability when exposed to 2,4-dinitrophenol. This suggests that adenosine triphosphate production in dystrophic muscle is partially inhibited as has been suggested by other workers.  相似文献   

15.
Female (I/St X C57BL/St) F1 mice heterozygous at the sex-linked phosphorylase kinase deficiency locus (Phk) have phosphorylase kinase activities averaging 86% that of mice homozygous for the wild-type allele (C57BL/St), i.e., 72% greater than the sum of one-half the activities of the parental strains. Approximately one-half the phosphorylase kinase activity in the (I X C57BL) F1 muscle extracts had a stability at 42.5 C similar to that of the activity in C57BL extracts (t1/2 = 13.2 min); the other half of the activity in the F1 extracts was more labile (t1/2 = 3.9 min). Two species of phosphorylase kinase activity in F1 muscle extracts were also differentiated with an antiserum prepared in guinea pigs against purified rabbit skeletal muscle phosphorylase kinase. This anti-serum cross-reacted with phosphorylase kinase in C57BL muscle extracts but did not cross-react with skeletal muscle extracts of mice hemi- or homozygous for the mutant allele (I/LnJ). The guinea pig antiserum precipitated 52% as much protein from (I X C57BL)F1 muscle extracts compared to those of C57BL. However, an antiserum prepared against purified rabbit skeletal muscle phosphorylase kinase in the goat cross-reacted with the mutant phosphorylase kinase. The ratio C57BL:(I X C57BL)F1:I of immunoprecipitated protein from skeletal muscle extracts with this antiserum was 1:0.97:1.08. Polyacrylamide gel electrophoresis of the immunoprecipitates in the presence of 0.1% sodium dodecylsulfate showed three subunits for mouse phosphorylase kinase with molecular weights of 139,000, 118,000, and 41,000; these values are similar to the ones obtained with purified rabbit skeletal muscle phosphorylase kinase. These three subunits were also observed in immunoprecipitates from I/LnJ muscle extracts. These results offer substantial evidence (1) that in skeletal muscle extracts of mice heterozygous at the Phk locus the mutant phosphorylase kinase is active, (2) that the gene product of the mutant allele is an enzyme with an abnormal structure, and (3) that the phosphorylase kinase deficiency in I/LnJ skeletal muscle extracts is not the result of the absence of phosphorylase kinase or one of its subunits.  相似文献   

16.
Previous studies have shown that norepinephrine (NE) elicits trans-membrane potential changes in skeletal muscle cells from normal and dystrophic (BIO 14.6) hamsters, with the magnitude of these changes being significantly less in dystrophic cells. To determine if the decreased response of the dystrophic muscle cells reflects a more generalized phenomenon, the present study was designed to evaluate the effects of NE on membrane properties of brown adipocytes. In vivo techniques using glass microelectrodes were similar to those used in the muscle studies. NE injection (2 to 5 μg/kg body wt, i.v.) into anesthetized hamsters was followed by membrane depolarization, the magnitude of which did not significantly differ in the dystrophic and normal adipocytes. For example, upon administration of 5 μg NE/kg body wt, the average depolarization was 14.5 ± 1.3 mV (X ± S.E.) for 20 dystrophic cells and 14.1 ± 1.8 mV for 18 normal cells. The depolarizations following i.v. infusion of isoproterenol and phenylephrine also had similar amplitudes in both normal and dystrophic cells. Despite this lack of difference in plasma membrane responses, NE induced a significantly smaller rise in interscapular brown fat temperature in the dystrophic (0.09°C) than in the normal hamsters (0.26°C) following administration of 5 μg NE/kg body wt. Thus, the decreased responsiveness to NE of dystrophic sarcolemma did not occur with the plasma membrane of brown adipocytes, although brown fat temperature changes in the dystrophic hamsters were decreased in amplitude.  相似文献   

17.
Mitochondria from skeletal muscle, heart and liver of strain 129/ReJ-dy dystrophic mice and their littermate controls were characterized with respect to their respiratory and phosphorylating activities. Skeletal muscle mitochondria from dystrophic mice showed significantly lower state 3 respiratory rates than controls with both pyruvate + malate and succinate as substrates (P < 0.01). ADP/O and Ca2+/O ratios were found to be normal. A decreased rate of NADH oxidation (0.01 <P < 0.05) by sonicated mitochondrial suspensions from dystrophic mice was also seen. High respiratory rates with ascorbate + phenazine methosulfate as substrates indicated that cytochrome oxidase was not rate limiting in the oxidation of either pyruvate + malate or succinate. Skeletal muscle mitochondria from dystrophic mice showed no deficiency in any of the cytochromes or coenzyme Q. Mg2+-stimulated ATPase activity was higher in dystrophic muscle mitochondria than in controls, but basal and oligomycin-insensitive activities were virtually identical to those of controls. A significant reduction in the intramitochondrial NAD+ content (0.01 <P < 0.02) was seen in dystrophic skeletal muscle as compared to controls. Heart mitochondria from dystrophic mice showed similar, though less extensive abnormalities while liver mitochondria were essentially normal. We concluded from these results that skeletal muscle mitochondria from strain 129 dystrophic mice possess impairments in substrate utilization which may result from (1) an abnormality in the transfer of electrons on the substrate side of coenzyme Q in the case of succinate oxidation; (2) a defect on the path of electron flow from NADH to cytochrome c, and (3) a deficiency of NAD+ in the case of NAD+-linked substrates.  相似文献   

18.
Inherited muscular dystrophy of the chicken is thought to arise from abnormal development of trophic regulation of skeletal muscles by their innervating nerves. To determine whether expression of muscular dystrophy in the chicken is a property of the nerves or of the muscles, wing limb buds were transplanted between normal and dystrophic chick embryos at 312 days of incubation (stage 19–20). Muscles of donor limbs innervated by nerves of the hosts were compared to contralateral unoperated host limb muscles in chicks from 6 to 25 weeks after hatching. Expression of normal or dystrophic phenotype was determined by examination of five different properties which are altered in dystrophic chick muscle: electromyographic evidence of myotonia; fiber diameter; acetylcholinesterase activity, localization, and isozymes; lactic dehydrogenase activity; and succinic dehydrogenase activity. Genetically normal muscle innervated by nerves of normal or dystrophic hosts was phenotypically normal while genetically dystrophic muscle innervated by normal nerves was phenotypically dystrophic. The results suggest that inherited muscular dystrophy of the chicken arises from a defect of muscle rather than from a lesion in the nerves themselves.  相似文献   

19.
The ouabain-sensitive (Na+ + K+)-ATPase activities of membrane fractions from hind-leg muscle and brain of normal and genetically dystrophic mice (C57BL/6J-dy strain) were studied, and the following results were obtained. 1) The ouabain-sensitive (Na+ + K+)-ATPase activity of frozen muscle sarcolemmal fraction from normal mice was several times higher than that of fresh one. 2) The ouabain-sensitive (Na+ + K+)-ATPase activity of frozen muscle sarcolemmal fraction from dystrophic mice was almost equal to that from normal one. But the muscle membrane yield from dystrophic mice was considerably low compared with the yield from normal one. 3) With brain membrane fractions, no differences were observed between normal and dystrophic mice in the ouabain-sensitive (Na+ + K+)-ATPase activity as well as in the yield of membrane fractions.  相似文献   

20.
The subcellular distribution of glycogen phosphorylase in pectoralis muscle from normal and dystrophic chickens was determined as a function of age. A substantially larger proportion of the total activity was associated with membranes cellular organelles, both mitochondria and sarcoplasmic reticulum, in preparations from dystrophic birds. The difference could be detected as early as 2 weeks ex ovo. Interaction of phosphorylase with cellular membranes may provide a probe for the underlying membrane defect in this dystrophyl model.  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