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 0 毫秒
1.
A cDNA library was constructed from mouse cardiac muscle mRNA, and a clone corresponding to part of the mRNA for the cardiac muscle isoform of actin was isolated from this library. The nucleotide sequence of the cloned insert was determined and was found to contain almost the complete amino acid coding region for actin (only codons for the first two amino acids, absent from the mature protein, were lacking) and a substantial portion derived from the 3 untranslated region of the mRNA. Comparison of the latter with the corresponding region in cardiac actin mRNA from man and rat showed that this 3 untranslated region has been subject to conservational pressure during evolution. However a comparison with the corresponding region in skeletal muscle actin mRNAs indicated that the pattern of conservation is quite different in the two striated muscle actin isoforms.  相似文献   

2.
The actins are a group of highly conserved proteins encoded by a multigene family. We have previously reported that the skeletal muscle actin gene is located on mouse chromosome 3, together with several other unidentified actin DNA sequences. We show here that the gene coding for the cardiac muscle actin, which is closely related to the skeletal muscle actin (1.1% amino acid replacements), is located on mouse chromosome 17. The gene coding for the cytoplasmic beta-actin is located on mouse chromosome 5. Thus, these three actin genes are located on three different chromosomes.  相似文献   

3.
Previous studies have shown that depletion of cardiac actin by targeted disruption is associated with increased expression of alternative actins in the mouse heart. Here we have studied the effects of transgenic overexpression of cardiac actin using the -myosin heavy chain promoter. Lines carrying 7 or 8 copies of the transgene showed a 2-fold increase in cardiac actin mRNA and also displayed decreased expression of skeletal and vascular actin in their hearts. In contrast, a line with more than 250 copies of the transgene did not show a similar decrease in the expression of skeletal and vascular actin despite a 3-fold increase in cardiac actin mRNA. While the low copy number transgenic mice displayed hearts that were similar to non-transgenic controls, the high copy number transgenic line showed larger hearts with distinct atrial enlargement and cardiomyocyte hypertrophy. Further, while the low copy number transgenic mouse hearts were mildly hypocontractile when compared with non-transgenic mouse hearts, the high copy number transgenic mouse hearts were significantly so. We conclude that in the presence of a small number of copies of the cardiac actin transgene, homeostatic mechanisms involved in maintaining actin levels are active and negatively regulate skeletal and vascular actin levels in the heart in response to increased expression of cardiac actin. However, these putative mechanisms are either inoperative in the high copy number transgenic line or are countered by the enhanced expression of skeletal and vascular actin during cardiomyocyte hypertrophy.  相似文献   

4.
5.
We have previously shown that actin from Tetrahymena pyriformis has a very divergent primary structure (Hirono, M., Endoh, H., Okada, N., Numata, O., & Watanabe, Y. (1987) J. Mol. Biol. 194, 181-192) and that though it shares essential properties with skeletal muscle actin, it does not interact at all with phalloidin or DNase I (Hirono, M., Kumagai, Y., Numata, O., & Watanabe, Y. (1989) Proc. Natl. Acad. Sci. U.S. 86, 75-79). In this study, we investigated the copolymerization of this actin with skeletal muscle actin by direct observation of the heteropolymers formed from the two actins by means of electron microscopy. We also examined the binding of actin-binding proteins from skeletal muscle or smooth muscle to Tetrahymena actin by means of a cosedimentation assay. The results show that (i) Tetrahymena actin copolymerizes with skeletal muscle actin and that (ii) muscle myosin subfragment 1 binds to it in the absence of ATP, like skeletal muscle actin. However, it was also shown that (iii) muscle alpha-actinin hardly binds to Tetrahymena actin and that (iv) muscle tropomyosin does not bind to it at all. The results show that Tetrahymena actin has both properties similar and dissimilar to those of skeletal muscle actin.  相似文献   

6.
7.
Salmonids utilize a unique, class II isoactin in slow skeletal muscle. This actin contains 12 replacements when compared with those from salmonid fast skeletal muscle, salmonid cardiac muscle and rabbit skeletal muscle. Substitutions are confined to subdomains 1 and 3, and most occur after residue 100. Depending on the pairing, the 'fast', 'cardiac' and rabbit actins share four, or fewer, substitutions. The two salmonid skeletal actins differ nonconservatively at six positions, residues 103, 155, 278, 281, 310 and 360, the latter involving a change in charge. The heterogeneity has altered the biochemical properties of the molecule. Slow skeletal muscle actin can be distinguished on the basis of mass, hydroxylamine cleavage and electrophoretic mobility at alkaline pH in the presence of 8 m urea. Further, compared with its counterpart in fast muscle, slow muscle actin displays lower activation of myosin in the presence of regulatory proteins, and weakened affinity for nucleotide. It is also less resistant to urea- and heat-induced denaturation. The midpoints of the change in far-UV ellipticity of G-actin versus temperature are approximately 45 degrees C ('slow' actin) and approximately 56 degrees C ('fast' actin). Similar melting temperatures are observed when thermal unfolding is monitored in the aromatic region, and is suggestive of differential stability within subdomain 1. The changes in nucleotide affinity and stability correlate with substitutions at the nucleotide binding cleft (residue 155), and in the C-terminal region, two parts of actin which are allosterically coupled. Actin is concluded to be a source of skeletal muscle plasticity.  相似文献   

8.
Among six actin isoforms, α-skeletal and α-cardiac actins have similar amino acid components and are highly conserved. Although skeletal muscles essentially express α-skeletal actins in the adult tissue, α-cardiac isoform actin is prominent in the embryonic muscle tissue. Switching of actin isoforms from α-cardiac to α-skeletal actin occurs during skeletal muscle differentiation. The cardiac type α-actin is expressed in the regeneration and patho-physiological states of the skeletal muscles as well. In the present study, we demonstrate the morphological switching of α-type actin isoforms from α-cardiac to α-skeletal actin in vitro using mouse ES cells for the first time. Immunofluorescent double staining with two specific antibodies revealed that α-cardiac actin appeared first in myoblasts. After cell fusion to form myotubes, the cardiac type actin decreased and α-skeletal actin conversely increased. Finally, the α-skeletal isoform remained as a main actin component in the fully mature skeletal muscle fibers. The exchange of isoforms is not directly linked to the sarcomere formation. As a result, ES cells provide a useful in vitro system for exploring skeletal muscle differentiation.  相似文献   

9.
Chloride activity and its control in skeletal and cardiac muscle   总被引:9,自引:0,他引:9  
Ion-selective microelectrodes have been used to compare the mechanisms controlling intracellular Cl- activity in skeletal and cardiac muscle. In frog Sartorius skeletal muscle fibres, Cl- levels are low (about 3 mM) and are determined mainly passively. The effect of any Cl- transport system will be quickly short-circuited through the high membrane Cl- conductance. In contrast, the sheep-heart Purkinje fibre, like other cardiac tissues, contains higher than passive levels of intracellular Cl- (20-30 mM). Many Cl- movements occur, not through Cl- channels (the permeability for Cl- is low), but by a Cl- -HCO3- countertransport system. High internal Cl- levels are achieved by an exchange of extracellular Cl- for intracellular HCO3-, which acidifies the fibre by 0.3 pH. Anion exchange in heart differs from that proposed for other excitable cells in that it is not specialized to compensate for an intracellular acidosis. Instead, it can prevent the fibres from becoming too alkaline by promoting a bicarbonate efflux and a chloride influx whenever internal bicarbonate levels rise. Possible reasons for this are briefly discussed.  相似文献   

10.
11.
The ability of actin to interact with hemin was studied. It was found that the Soret absorption band of hemin changes in the presence of actin and that hemin is capable of quenching the fluorescence intensity of actin. These findings were indicative of hemin binding to actin. The binding constant for the high affinity site was calculated to be 5.3 X 10(6) M-1. The amounts of native G- and F-actin were estimated by their DNAase I inhibition activity. It was observed that the binding of hemin to G-actin is followed by a slow decrease in the ability of actin to inhibit DNAase I activity and to polymerize upon addition of salts. Binding of hemin to F-actin resulted in a gradual depolymerization of the filaments, to an inactivated form, as expressed by a reduction in the ability of hemin-bound F-actin to inhibit DNAase I activity in the absence as well as in the presence of guanidine-HCl. Electron microscopy studies further corroborated these findings by demonstrating that: (1) hemin-bound G-actin failed to show formation of polymers when salts were added; (2) a marked reduction in the amount of actin polymers was observed in the specimens examined 24 h after mixing with hemin. It is suggested that the elevated amounts of free hemin formed under pathological conditions, might be toxic to cells by interfering with actin polymerization cycles.  相似文献   

12.
13.
Actin isoform expression may change during development, and in certain physiological, experimental and pathological situations. It is accepted that during sarcomeric (skeletal and cardiac) muscle development, the alpha-skeletal and alpha-cardiac isoforms of actin accumulate rapidly at the onset of muscle fibre formation, while there is a rapid fall in the expression of nonmuscle (beta and gamma) actin isoforms. Here we show that, before birth, both skeletal and myocardial cells express significant amounts of alpha-smooth muscle actin mRNA and protein. This expression is transient and disappears over the 1-7 days following birth. Our findings show that the program regulating actin isoform expression in sarcomeric muscle development is complex and that alpha-smooth muscle actin participates in this process.  相似文献   

14.
15.
1. Mouse ventricle strips provide a good model system for studying cellular damage in mammalian cardiac muscle. 2. Anoxia rapidly causes destruction of the myofilament apparatus that is characteristic of calcium-triggered damage in muscle cells, and it is suggested that anoxia promotes release of calcium from the mitochondria. 3. Oxygen exacerbates this damage which is independent of extracellular calcium; it is suggested that it initiates myofilament damage by activation at an intracellular site, probably the sarcoplasmic reticulum.  相似文献   

16.
SMAD4 acts as the converging point for TGFβ and BMP signaling in heart development. Here, we investigated the role of SMAD4 in heart development using a novel α skeletal muscle actin Cre recombinase (MuCre) transgenic mouse strain. Lineage tracing using MuCre/ROSA26LacZ reporter mice indicated strong Cre-recombinase expression in developing and adult heart and skeletal muscles. In heart development, significant MuCre expression was noted at E11.5 in the atrial, ventricular, outflow tract and atrioventricular canal myocardium, but not in the endocardial cushions. MuCre-driven conditional deletion of Smad4 in mice caused double outlet right ventricle (DORV), ventricular septal defect (VSD), impaired trabeculation and thinning of ventricular myocardium, and mid-gestational embryonic lethality. In conclusion, MuCre mice effectively delete genes in both heart and skeletal muscles, thus enabling the discovery that myocardial Smad4 deletion causes misalignment of the outflow tract and DORV.  相似文献   

17.
18.
We have recently reported that actin modified with dimethyl suberimidate takes a filamentous form even under depolymerizing conditions, and this phenomenon is accounted for by the conformational fixation caused by the introduction of an intramolecular cross-link (Ohara, O., Takahashi, S., Ooi, T., & Fujiyoshi, Y. (1982) J. Biochem. 91, 1999-2012). The suberimidate-treated actin (SA) is not immediately depolymerized by deoxyribonuclease I (DNase I) but is depolymerized after incubation for one day, i.e., depolymerization is much slower than that for intact F-actin. The results on circular dichroic spectra of a mixture of SA and DNase I suggest that DNase I flips the conformation of SA into a G-actin-like state from the F-actin-like one when a tight SA-DNase I complex is formed. The suberimidate cross-link introduced in an SA molecule does not completely prevent the conformational change from the F-state to the G-state but stabilizes the actin conformation very greatly in the F-state.  相似文献   

19.
Several ganglioside-binding proteins have been identified inguinea pig skeletal and cardiac muscle. In the cytosolic fractionsof both tissues, a 130-kD protein was found to have the highestpropensity to bind lucifer yellow CH-labelled GM1. This bindingcould be abolished by prior incubation of the protein with GM2.Polysialogangliosides including GD1a, GD1b, GT1b, and GQ1b wereless effective. The 130-kD protein migrated as a doublet withapparent isoelectric points (pI) of 6.3 and 6.5, respectively,in isoelectric focusing gel, but as a single species with anapparent Mr of 43000 in SDS-polyacrylamide gel. Both the ganglioside-bindingand the immunological properties of the 43-kD subunit proteinwere different from those of rabbit skeletal muscle actin. Cardiacmuscle extract also contained a 77-kD minor ganglioside-bindingprotein that was absent in skeletal muscle. This protein hadan apparent pI of 5.4 and migrated as a 39-kD species in SDSgels. By contrast, only the particulate fraction of skeletalmuscle was found to contain a 180-kD major ganglioside-bindingprotein. Binding of fluorescent GM1 to this protein was blockedby pre-incubation of the protein with GM1 or GM2. The 180-kDprotein migrated as a 98-kD species in SDS gels. However, itspropensity to bind lucifer yellow CH-GM1 was at least 10 timesgreater than that of rabbit skeletal muscle phosphorylase b(Mr = 97400). The apparent pI (6.5) of the 180-kD protein alsowas slightly higher than that of rabbit phosphorylase. Tissuedistribution studies revealed that both the 130-kD and the 180-kDmajor ganglioside-binding proteins were muscle specific. Itis, therefore, possible that these two proteins may play someunique roles in ganglioside-related functions in muscle tissues. gangliosides ganglioside-binding proteins muscle  相似文献   

20.
MicroRNAs in skeletal and cardiac muscle development   总被引:1,自引:0,他引:1  
MicroRNAs (miRNAs) are a recently discovered class of small non-coding RNAs, which are approximately 22 nucleotides in length. miRNAs negatively regulate gene expression by translational repression and target mRNA degradation. It has become clear that miRNAs are involved in many biological processes, including development, differentiation, proliferation, and apoptosis. Interestingly, many miRNAs are expressed in a tissue-specific manner and several miRNAs are specifically expressed in cardiac and skeletal muscles. In this review, we focus on those miRNAs that have been shown to be involved in muscle development. Compelling evidences have demonstrated that muscle miRNAs play an important role in the regulation of muscle proliferation and differentiation processes. However, it appears that miRNAs are not essential for early myogenesis and muscle specification. Importantly, dysregulation of miRNAs has been linked to muscle-related diseases, such as cardiac hypertrophy. A mutation resulting in a gain-of-function miRNA target site in the myostatin gene leads to down regulation of the targeted protein in Texel sheep. miRNAs therefore are a new class of regulators of muscle biology and they might become novel therapeutic targets in muscle-related human diseases.  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