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 15 毫秒
1.
Phosphorylation on Ser 19 of the myosin II regulatory light chain by myosin light chain kinase (MLCK) regulates actomyosin contractility in smooth muscle and vertebrate nonmuscle cells. The smooth/nonmuscle MLCK gene locus produces two kinases, a high molecular weight isoform (long MLCK) and a low molecular weight isoform (short MLCK), that are differentially expressed in smooth and nonmuscle tissues. To study the relative localization of the MLCK isoforms in cultured nonmuscle cells and to determine the spatial and temporal dynamics of MLCK localization during mitosis, we constructed green fluorescent protein fusions of the long and short MLCKs. In interphase cells, localization of the long MLCK to stress fibers is mediated by five DXRXXL motifs, which span the junction of the NH(2)-terminal extension and the short MLCK. In contrast, localization of the long MLCK to the cleavage furrow in dividing cells requires the five DXRXXL motifs as well as additional amino acid sequences present in the NH(2)-terminal extension. Thus, it appears that nonmuscle cells utilize different mechanisms for targeting the long MLCK to actomyosin structures during interphase and mitosis. Further studies have shown that the long MLCK has twofold lower kinase activity in early mitosis than in interphase or in the early stages of postmitotic spreading. These findings suggest a model in which MLCK and the myosin II phosphatase (Totsukawa, G., Y. Yamakita, S. Yamashiro, H. Hosoya, D.J. Hartshorne, and F. Matsumura. 1999. J. Cell Biol. 144:735-744) act cooperatively to regulate the level of Ser 19-phosphorylated myosin II during mitosis and initiate cytokinesis through the activation of myosin II motor activity.  相似文献   

2.
3.
4.
The Ca(2+)/calmodulin-dependent endothelial cell myosin light chain kinase (MLCK) triggers actomyosin contraction essential for vascular barrier regulation and leukocyte diapedesis. Two high molecular weight MLCK splice variants, EC MLCK-1 and EC MLCK-2 (210-214 kDa), in human endothelium are identical except for a deleted single exon in MLCK-2 encoding a 69-amino acid stretch (amino acids 436-505) that contains potentially important consensus sites for phosphorylation by p60(Src) kinase (Lazar, V., and Garcia, J. G. (1999) Genomics 57, 256-267). We have now found that both recombinant EC MLCK splice variants exhibit comparable enzymatic activities but a 2-fold reduction of V(max), and a 2-fold increase in K(0.5 CaM) when compared with the SM MLCK isoform, whereas K(m) was similar in the three isoforms. However, only EC MLCK-1 is readily phosphorylated by purified p60(Src) in vitro, resulting in a 2- to 3-fold increase in EC MLCK-1 enzymatic activity (compared with EC MLCK-2 and SM MLCK). This increased activity of phospho-MLCK-1 was observed over a broad range of submaximal [Ca(2+)] levels with comparable EC(50) [Ca(2+)] for both phosphorylated and unphosphorylated EC MLCK-1. The sites of tyrosine phosphorylation catalyzed by p60(Src) are Tyr(464) and Tyr(471) within the 69-residue stretch deleted in the MLCK-2 splice variant. These results demonstrate for the first time that p60(Src)-mediated tyrosine phosphorylation represents an important mechanism for splice variant-specific regulation of nonmuscle MLCK and vascular cell function.  相似文献   

5.
The myosin 20,000-D regulatory light chain (RLC) has a central role in smooth muscle contraction. Previous work has suggested either the presence of two RLC isoforms, one specific for nonmuscle and one specific for smooth muscle, or the absence of a true smooth muscle-specific isoform, in which instance smooth muscle cells would use nonmuscle isoforms. To address this issue directly, we have isolated rat RLC cDNAs and corresponding genomic sequences of two smooth muscle RLC based on homology to the amino acid sequence of the chicken gizzard RLC. These cDNAs are highly homologous in their amino acid coding regions and contain unique 3'-untranslated regions. RNA analyses of rat tissue using these unique 3'-untranslated regions revealed that their expression is differentially regulated. However, one cDNA (RLC-B), predominantly a nonmuscle isoform, based on abundant expression in nonmuscle tissues including brain, spleen, and lung, is easily detected in smooth muscle tissues. The other cDNA (RLC-A; see Taubman, M., J. W. Grant, and B. Nadal-Ginard. 1987. J. Cell Biol. 104:1505-1513) was detected in a variety of nonmuscle, smooth muscle, and sarcomeric tissues. RNA analyses comparing expression of both RLC genes with the actin gene family and smooth muscle specific alpha-tropomyosin demonstrated that neither RLC gene was strictly smooth muscle specific. RNA analyses of cell lines demonstrated that both of the RLC genes are expressed in a variety of cell types. The complete genomic structure of RLC-A and close linkage to RLC-B is described.  相似文献   

6.
The purpose of this study was to characterize myosin light chain kinase (MLCK) expression in cardiac and skeletal muscle. The only classic MLCK detected in cardiac tissue, purified cardiac myocytes, and in a cardiac myocyte cell line (AT1) was identical to the 130-kDa smooth muscle MLCK (smMLCK). A complex pattern of MLCK expression was observed during differentiation of skeletal muscle in which the 220-kDa-long or "nonmuscle" form of MLCK is expressed in undifferentiated myoblasts. Subsequently, during myoblast differentiation, expression of the 220-kDa MLCK declines and expression of this form is replaced by the 130-kDa smMLCK and a skeletal muscle-specific isoform, skMLCK in adult skeletal muscle. These results demonstrate that the skMLCK is the only tissue-specific MLCK, being expressed in adult skeletal muscle but not in cardiac, smooth, or nonmuscle tissues. In contrast, the 130-kDa smMLCK is ubiquitous in all adult tissues, including skeletal and cardiac muscle, demonstrating that, although the 130-kDa smMLCK is expressed at highest levels in smooth muscle tissues, it is not a smooth muscle-specific protein.  相似文献   

7.
We have isolated tropomyosin cDNAs from human skeletal muscle and nonmuscle cDNA libraries and constructed gene-specific DNA probes for each of the four functional tropomyosin genes. These DNA probes were used to define the regulation of the corresponding mRNAs during the process of myogenesis. Tropomyosin regulation was compared with that of beta- and gamma-actin. No two striated muscle-specific tropomyosin mRNAs are coordinately accumulated during myogenesis nor in adult striated muscles. Similarly, no two nonmuscle tropomyosins are coordinately repressed during myogenesis. However, mRNAs encoding the 248 amino acid nonmuscle tropomyosins and beta- and gamma-actin are more persistent in adult skeletal muscle than those encoding the 284 amino acid nonmuscle tropomyosins. In particular, the nonmuscle tropomyosin Tm4 is expressed at similar levels in adult rat nonmuscle and striated muscle tissues. We conclude that each tropomyosin mRNA has its own unique determinants of accumulation and that the 248 amino acid nonmuscle tropomyosins may have a role in the architecture of the adult myofiber. The variable regulation of nonmuscle isoforms during myogenesis suggests that the different isoforms compete for inclusion into cellular structures and that compensating autoregulation of mRNA levels bring gene expression into alignment with the competitiveness of each individual gene product. Such an isoform competition-autoregulatory compensation mechanism would readily explain the unique regulation of each gene.  相似文献   

8.
We have investigated accumulation of alpha-actinin, the principal cross-linker of actin filaments, in four Drosophila fliA mutants. A single gene is variably spliced to generate one nonmuscle and two muscle isoforms whose primary sequence differences are confined to a peptide spanning the actin binding domain and first central repeat. In fliA3 the synthesis of an adult muscle-specific isoform is blocked in flight and leg muscles, while in fliA4 the synthesis of nonmuscle and both muscle-specific isoforms is severely reduced. Affected muscles are weak or paralyzed, and, in the case of fliA3, atrophic. Their myofibrils, while structurally irregular, are remarkably normal considering that they are nearly devoid of a major contractile protein. Also surprising is that no obvious nonmuscle cell abnormalities can be discerned despite the fact that both the fliA1- and fliA4-associated mutations perturb the nonmuscle isoform. Our observations suggest that alpha-actinin stabilizes and anchors thin filament arrays, rather than orchestrating their assembly, and further imply that alpha-actinin function is redundant in both muscle and nonmuscle cells.  相似文献   

9.
We report that the genetic locus that encodes vertebrate smooth muscle and nonmuscle myosin light chain kinase (MLCK) and kinase-related protein (KRP) has a complex arrangement and a complex pattern of expression. Three proteins are encoded by 31 exons that have only one variation, that of the first exon of KRP, and the genomic locus spans approximately 100 kb of DNA. The three proteins can differ in their relative abundance and localization among tissues and with development. MLCK is a calmodulin (CaM) regulated protein kinase that phosphorylates the light chain of myosin II. The chicken has two MLCK isoforms encoded by the MLCK/KRP locus. KRP does not bind CaM and is not a protein kinase. However, KRP binds to and regulates the structure of myosin II. Thus, KRP and MLCK have the same subcellular target, the myosin II molecular motor system. We examined the tissue and cellular localization of KRP and MLCK in the chicken embryo and in adult chicken tissues. We report on the selective localization of KRP and MLCK among and within tissues and on a differential distribution of the proteins between embryonic and adult tissues. The results fill a void in our knowledge about the organization of the MLCK/KRP genetic locus, which appears to be a late evolving regulatory paradigm, and suggest an independent and complex regulation of expression of the gene products from the MLCK/KRP genetic locus that may reflect a basic principle found in other eukaryotic gene clusters that encode functionally linked proteins. J. Cell. Biochem. 70:402–413, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
To better understand thedistinct functional roles of the 220- and 130-kDa forms of myosin lightchain kinase (MLCK), expression and intracellular localization weredetermined during development and in adult mouse tissues. Northernblot, Western blot, and histochemical studies show that the 220-kDaMLCK is widely expressed during development as well as in several adultsmooth muscle and nonmuscle tissues. The 130-kDa MLCK is highlyexpressed in all adult tissues examined and is also detectable duringembryonic development. Colocalization studies examining thedistribution of 130- and 220-kDa mouse MLCKs revealed that the 130-kDaMLCK colocalizes with nonmuscle myosin IIA but not with myosin IIB orF-actin. In contrast, the 220-kDa MLCK did not colocalize with eithernonmuscle myosin II isoform but instead colocalizes with thickinterconnected bundles of F-actin. These results suggest that in vivo,the physiological functions of the 220- and 130-kDa MLCKs are likely tobe regulated by their intracellular trafficking and distribution.

  相似文献   

11.
Despite their near sequence identity, actin isoforms cannot completely replace each other in vivo and show marked differences in their tissue-specific and subcellular localization. Little is known about isoform-specific differences in their interactions with myosin motors and other actin-binding proteins. Mammalian cytoplasmic β- and γ-actin interact with nonsarcomeric conventional myosins such as the members of the nonmuscle myosin-2 family and myosin-7A. These interactions support a wide range of cellular processes including cytokinesis, maintenance of cell polarity, cell adhesion, migration, and mechano-electrical transduction. To elucidate differences in the ability of isoactins to bind and stimulate the enzymatic activity of individual myosin isoforms, we characterized the interactions of human skeletal muscle α-actin, cytoplasmic β-actin, and cytoplasmic γ-actin with human myosin-7A and nonmuscle myosins-2A, -2B and -2C1. In the case of nonmuscle myosins-2A and -2B, the interaction with either cytoplasmic actin isoform results in 4-fold greater stimulation of myosin ATPase activity than was observed in the presence of α-skeletal muscle actin. Nonmuscle myosin-2C1 is most potently activated by β-actin and myosin-7A by γ-actin. Our results indicate that β- and γ-actin isoforms contribute to the modulation of nonmuscle myosin-2 and myosin-7A activity and thereby to the spatial and temporal regulation of cytoskeletal dynamics. FRET-based analyses show efficient copolymerization abilities for the actin isoforms in vitro. Experiments with hybrid actin filaments show that the extent of actomyosin coupling efficiency can be regulated by the isoform composition of actin filaments.  相似文献   

12.
Muscle creatine kinase (MCK) is expressed at high levels only in skeletal and cardiac muscle tissues. Previous in vitro transfection studies of skeletal muscle myoblasts and fibroblasts had identified two MCK enhancer elements and one proximal promoter element, each of which exhibited expression only in differentiated skeletal muscle. In this study, we have identified several regions of the mouse MCK gene that are responsible for tissue-specific expression in transgenic mice. A fusion gene containing 3,300 nucleotides of MCK 5' sequence exhibited chloramphenicol acetyltransferase activity levels that were more than 10(4)-fold higher in skeletal muscle than in other, nonmuscle tissues such as kidney, liver, and spleen. Expression in cardiac muscle was also greater than in these nonmuscle tissues by 2 to 3 orders of magnitude. Progressive 5' deletions from nucleotide -3300 resulted in reduced expression of the transgene, and one of these resulted in a preferential decrease in expression in cardiac tissue relative to that in skeletal muscle. Of the two enhancer sequences analyzed, only one directed high-level expression in both skeletal and cardiac muscle. The other enhancer activated expression only in skeletal muscle. These data reveal a complex set of cis-acting sequences that have differential effects on MCK expression in skeletal and cardiac muscle.  相似文献   

13.
Activation of Na(+)-nutrient cotransport leads to increased tight junction permeability in intestinal absorptive (villus) enterocytes. This regulation requires myosin II regulatory light chain (MLC) phosphorylation mediated by MLC kinase (MLCK). We examined the spatiotemporal segregation of MLCK isoform function and expression along the crypt-villus axis and found that long MLCK, which is expressed as two alternatively spliced isoforms, accounts for 97 +/- 4% of MLC kinase activity in interphase intestinal epithelial cells. Expression of the MLCK1 isoform is limited to well differentiated enterocytes, both in vitro and in vivo, and this expression correlates closely with development of Na(+)-nutrient cotransport-dependent tight junction regulation. Consistent with this role, MLCK1 is localized to the perijunctional actomyosin ring. Furthermore, specific knockdown of MLCK1 using siRNA reduced tight junction permeability in monolayers with active Na(+)-glucose cotransport, confirming a functional role for MLCK1. These results demonstrate unique physiologically relevant patterns of expression and subcellular localization for long MLCK isoforms and show that MLCK1 is the isoform responsible for tight junction regulation in absorptive enterocytes.  相似文献   

14.
We have previously shown that chicken embryo fibroblast (CEF) cells and human bladder carcinoma (EJ) cells contain multiple isoforms of tropomyosin, identified as a, b, 1, 2, and 3 in CEF cells and 1, 2, 3, 4, and 5 in human EJ cells by one-dimensional SDS-PAGE (Lin, J. J.-C., D. M. Helfman, S. H. Hughes, and C.-S. Chou. 1985. J. Cell Biol. 100: 692-703; and Lin, J. J.-C., S. Yamashiro-Matsumura, and F. Matsumura. 1984. Cancer Cells 1:57-65). Both isoform 3 (TM-3) of CEF and isoforms 4,5 (TM-4,-5) of human EJ cells are the minor isoforms found respectively in normal chicken and human cells. They have a lower apparent molecular mass and show a weaker affinity to actin filaments when compared to the higher molecular mass isoforms. Using individual tropomyosin isoforms immobilized on nitrocellulose papers and sequential absorption of polyclonal antiserum on these papers, we have prepared antibodies specific to CEF TM-3 and to CEF TM-1,-2. In addition, two of our antitropomyosin mAbs, CG beta 6 and CG3, have now been demonstrated by Western blots, immunoprecipitation, and two-dimensional gel analysis to have specificities to human EJ TM-3 and TM-5, respectively. By using these isoform-specific reagents, we are able to compare the intracellular localizations of the lower and higher molecular mass isoforms in both CEF and human EJ cells. We have found that both lower and higher molecular mass isoforms of tropomyosin are localized along stress fibers of cells, as one would expect. However, the lower molecular mass isoforms are also distributed in regions near ruffling membranes. Further evidence for this different localization of different tropomyosin isoforms comes from double-label immunofluorescence microscopy on the same CEF cells with affinity-purified antibody against TM-3, and monoclonal CG beta 6 antibody against TM-a, -b, -1, and -2 of CEF tropomyosin. The presence of the lower molecular mass isoform of tropomyosin in ruffling membranes may indicate a novel way for the nonmuscle cell to control the stability and organization of microfilaments, and to regulate the cell motility.  相似文献   

15.
16.
Overlapping cDNA clones encoding a low M gamma human nonmuscle caldesmon isoform (HUM 1-CaD) span the entire coding region (538 amino acids) as well as 111 base pairs (bp) of 5'-noncoding and 1249 bp of 3'-noncoding region. Northern blot probes derived from either the coding or 3'-noncoding region hybridized to a 4.3-kilobase mRNA in nonmuscle cells and a 5.2-kilobase mRNA in stomach tissue. Primer extension results indicated that the 5'-noncoding region of the HUM 1-CaD mRNA is approximately 700 bp in length and also suggested that 1-CaD mRNAs with common 5'-noncoding regions are expressed in both liver and fibroblast cells. Comparisons of the human, rat, and chicken 1-CaD amino acids sequences demonstrated that although each isoform has unique characteristics, extensive regions of conservation exist. Amino acids 27-53 and 97-127 are 100% identical in these isoforms while amino acids 297-531 of HUM 1-CaD are 94 and 85% identical to the rat and chicken 1-CaDs, respectively. In addition, the levels of HUM 1-CaD mRNA and protein appeared to be decreased by 2-4 fold in the transformed derivatives of KD and WI38 cell lines as judged by Northern and Western blot analysis. The results suggest that the decrease of 1-CaD protein in these transformed cells is a direct result of decreased 1-CaD mRNA synthesis and/or increased mRNA turnover.  相似文献   

17.
Subunit VIIa of mammalian cytochrome c oxidase (COX; EC 1.9.3.1) exists in at least two isoforms, one present in all tissue types ('liver' isoform; COX VIIa-L) and the other specific for cardiac and skeletal muscle (COX VIIa-M). We have isolated a full-length cDNA encoding human COX VIIa-M. The deduced polypeptide represents the human ortholog of COX VIIa-M, as it shares 78% identity with bovine COX VIIa-M, but only 63% identity with human COX VIIa-L. Northern-blot analysis of primate tissues demonstrated that COXVIIa-M mRNA is present only in muscle tissues; in contrast, the COXVIIa-L mRNA is present in both muscle and nonmuscle tissues. Southern-blot hybridization of human-rodent cell hybrid genomic DNA indicates that the COXVIIa-M gene maps to a single locus on chromosome 19, designated COX7AM. In contrast, COXVIIa-L cDNA probes hybridized to fragments from two COX7AL loci, on chromosomes 4 and 14.  相似文献   

18.
19.
Isoproteins of myosin alkali light chain (LC) were co-expressed in cultured chicken cardiomyocytes and fibroblasts and their incorporation levels into myofibrils and stress fibers were compared among members of the LC isoform family. In order to distinguish each isoform from the other, cDNAs of LC isoforms were tagged with different epitopes. Expressed LCs were detected with antibodies to the tags and their distribution was analyzed by confocal microscopy. In cardiomyocytes, the incorporation level of LC into myofibrils was shown to increase in the order from nonmuscle isoform (LC3nm), to slow skeletal muscle isoform (LC1sa), to slow skeletal/ventricular muscle isoform (LC1sb), and to fast skeletal muscle isoforms (LC1f and LC3f). Thus, the hierarchal order of the LC affinity for the cardiac myosin heavy chain (MHC) is identical to that obtained in the rat (Komiyama et al., 1996. J. Cell Sci., 109: 2089-2099), suggesting that this order may be common for taxonomic animal classes. In fibroblasts, the affinity of LC for the nonmuscle MHC in stress fibers was found to increase in the order from LC3nm, to LC1sb, to LC1sa, and to LC1f and LC3f. This order for the nonmuscle MHC is partly different from that for the cardiac MHC. This indicates that the order of the affinity of LC isoproteins for MHC varies depending on the MHC isoform. Further, for both the cardiac and nonmuscle MHCs, the fast skeletal muscle LCs exhibited the highest affinity. This suggests that the fast skeletal muscle LCs may be evolved isoforms possessing the ability to associate tightly with a variety of MHC isoforms.  相似文献   

20.
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