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.
Myosin heavy chains (MHCs) from rat aorta smooth muscle cells were analyzed prior to and after these cells were placed into cell culture using sodium dodecyl sulfate-5% polyacrylamide gels, immunoblots, and two-dimensional peptide maps of tryptic digests. Rat aorta smooth muscle cells prior to culture were found to contain two MHCs (mass = 204 and 200 kDa) which cross-reacted with antibodies raised to smooth muscle myosin, but not with antibodies raised to platelet myosin. Tryptic peptide maps of these two MHCs showed no major differences when compared to each other and to maps of vas deferens and uterus smooth muscle MHCs. When rat aorta smooth muscle cells were placed into culture, the MHCs isolated from the cell extracts differed, depending on whether the cells were rapidly growing or postconfluent. Extracts from log-phase cultures contained predominantly MHCs that migrated more rapidly than smooth muscle myosin in sodium dodecyl sulfate-polyacrylamide gel electrophoresis (mass = 196 kDa) and cross-reacted with antibodies raised to platelet myosin, but not to smooth muscle myosin. Tryptic peptide maps of this MHC were very similar to those obtained with MHCs from non-muscle sources such as platelets and fibroblasts. In contrast, extracts from postconfluent rat aorta cell cultures contained three MHCs (mass = 204, 200, and 196 kDa). Using immunoblots and peptide maps, the fastest migrating MHC was found to be identical to the 196-kDa non-muscle MHC, while the two slower migrating MHCs had the same properties as aorta smooth muscle MHCs prior to culture. These results suggest that smooth muscle cells grown in primary culture contain predominantly (greater than 80%) non-muscle myosin while actively growing, but at a postconfluent stage, contain more equivalent amounts of smooth muscle and non-muscle myosins.  相似文献   

2.
We explored the hypothesis that discrepancies in the literature concerning the nature of myosin expression in cultured smooth muscle cells are due to the appearance of a new form of myosin heavy chain (MHC) in vitro. Previously, we used a very porous sodium dodecyl sulfate gel electrophoresis system to detect two MHCs in intact smooth muscles (SM1 and SM2) which differ by less than 2% in molecular weight (Rovner, A. S., Thompson, M. M., and Murphy, R. A. (1986) Am. J. Physiol. 250, C861-C870). Myosin-containing homogenates of rat aorta cells in primary culture were electrophoresed on this gel system, and Western blots were performed using smooth muscle-specific and nonmuscle-specific myosin antibodies. Subconfluent, rapidly proliferating cultures contained a form of heavy chain not found in rat aorta cells in vivo (NM) with electrophoretic mobility and antigenicity identical to the single unique heavy chain seen in nonmuscle cells. Moreover, these cultures expressed almost none of the smooth muscle heavy chains. In contrast, postconfluent growth-arrested cultures expressed increased levels of the two smooth muscle heavy chains, along with large amounts of NM. Analysis of cultures pulsed with [35S] methionine indicated that subconfluent cells were synthesizing almost exclusively NM, whereas postconfluent cells synthesized SM1 and SM2 as well as larger amounts of NM. Similar patterns of MHC content and synthesis were found in subconfluent and postconfluent passaged cells. These results show that cultured vascular smooth muscle cells undergo differential expression of smooth muscle- and nonmuscle-specific MHC forms with changes in their growth state, which appear to parallel changes in expression of the smooth muscle and nonmuscle forms of actin (Owens, G. K., Loeb, A., Gordon, D., and Thompson, M. M. (1986) J. Cell Biol. 102, 343-352). The reappearance of the smooth muscle MHCs in postconfluent cells suggests that density-related growth arrest promotes cytodifferentiation, but the continued expression of the nonmuscle MHC form in these smooth muscle cells indicates that other factors are required to induce the fully differentiated state while in culture.  相似文献   

3.
To describe phenotypic changes of human aortic smooth muscle cells (SMCs), proportion of smooth muscle and nonmuscle variants of actin, myosin heavy chains (MHCs), vinculin, and caldesmon, during prenatal and several months of postnatal development was determined. In aortic SMCs from 9-10-week-old fetus, both nonmuscle and smooth muscle-specific variants of all four proteins were present, however, the nonmuscle forms were more abundant. During development, a shift towards the expression of muscle-specific variants was observed, although the time course of changes in protein variant content was not similar for all the proteins studied. By the 24th week of gestation, fractional content of alpha-smooth muscle actin and smooth muscle MHCs was rather close to that in the mature SMCs, and comprised approximately 80 and 90%, respectively, of the levels characteristic of SMCs from adult aortic media. On the contrary, fractional ratio of meta-vinculin and 150-kDa caldesmon was still rather low in the aorta from the 24-week-old fetus, did not increase in a 2-month-old child aorta, and did not reach the level characteristic of mature SMCs even in the 6-month-old child aorta. Thus changes in alpha-smooth muscle actin and smooth muscle MHC fractional content occur mainly during the prenatal period of development, before the 24th week of gestation; while meta-vinculin and the 150-kDa caldesmon proportion increases mainly in the postnatal period, during several months after birth. In the "Discussion," phenotypes of SMCs from developing aorta were compared to those from different layers of the adult aortic wall.  相似文献   

4.
5.
The 20-kDa regulatory myosin light chain (MLC), also known as MLC-2, plays an important role in the regulation of both smooth muscle and nonmuscle cell contractile activity. Phosphorylation of MLC-2 by the enzyme MLC kinase increases the actin-activated myosin ATPase activity and thereby regulates the contractile activity. We have isolated and characterized an MLC-2 cDNA corresponding to the human vascular smooth muscle MLC-2 isoform from a cDNA library derived from umbilical artery RNA. The translation of the in vitro synthesized mRNA, corresponding to the cDNA insert, in a rabbit reticulocyte lysate results in the synthesis of a 20,000-dalton protein that is immunoreactive with antibodies raised against purified chicken gizzard MLC-2. The derived amino acid sequence of the putative human smooth muscle MLC-2 shows only three amino acid differences when compared to chicken gizzard MLC-2. However, comparison with the human cardiac isoform reveals only 48% homology. Blot hybridizations and S1 nuclease analysis indicate that the human smooth muscle MLC-2 isoform is expressed restrictively in smooth muscle tissues such as colon and uterus and in some, but not all, nonmuscle cell lines. Previously reported MLC-2 cDNA from rat aortic smooth muscle cells in culture is ubiquitously expressed in all muscle and nonmuscle cells, and it was suggested that both smooth muscle and nonmuscle MLC-2 proteins are identical and are probably encoded by the same gene. In contrast, the human smooth muscle MLC-2 cDNA that we have characterized from an intact smooth muscle tissue is not expressed in skeletal and cardiac muscles and also in a number of nonmuscle cells.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Maturation rates of vascular and visceral smooth muscle (SM) during ovine development were compared by quantifying contractile protein, myosin heavy chain (MHC) isoform contents, and contractile properties of aortas and bladders from female fetal (n = 19) and postnatal (n = 21) sheep. Actin, myosin, and protein contents rose progressively throughout development in both tissues (P 相似文献   

7.
Segregated assembly of muscle myosin expressed in nonmuscle cells.   总被引:6,自引:2,他引:4       下载免费PDF全文
Skeletal muscle myosin cDNAs were expressed in a simian kidney cell line (COS) and a mouse myogenic cell line to investigate the mechanisms controlling early stages of myosin filament assembly. An embryonic chicken muscle myosin heavy chain (MHC) cDNA was linked to constitutive promoters from adenovirus or SV40 and transiently expressed in COS cells. These cells accumulate hybrid myosin molecules composed of muscle MHCs and endogenous, nonmuscle, myosin light chains. The muscle myosin is found associated with a Triton insoluble fraction from extracts of the COS cells by immunoprecipitation and is detected in 2.4 +/- 0.8-micron-long filamentous structures distributed throughout the cytoplasm by immunofluorescence microscopy. These structures are shown by immunoelectron microscopy to correspond to loosely organized bundles of 12-16-nm-diameter myosin filaments. The muscle and nonmuscle MHCs are segregated in the transfected cells; the endogenous nonmuscle myosin displays a normal distribution pattern along stress fibers and does not colocalize with the muscle myosin filament bundles. A similar assembly pattern and distribution are observed for expression of the muscle MHC in a myogenic cell line. The myosin assembles into filament bundles, 1.5 +/- 0.6 micron in length, that are distributed throughout the cytoplasm of the undifferentiated myoblasts and segregated from the endogenous nonmuscle myosin. In both cell lines, formation of the myosin filament bundles is dependent on the accumulation of the protein. In contrast to these results, the expression of a truncated MHC that lacks much of the rod domain produces an assembly deficient molecule. The truncated MHC is diffusely distributed throughout the cytoplasm and not associated with cellular stress fibers. These results establish that the information necessary for the segregation of myosin isotypes into distinct cellular structures is contained within the primary structure of the MHC and that other factors are not required to establish this distribution.  相似文献   

8.
Treatment of human platelets with 162 nM 12-O-tetradecanoylphorbol-13-acetate (TPA) resulted in phosphorylation of a number of peptides, including myosin heavy chain and the 20-kDa myosin light chain. The site phosphorylated on the myosin heavy chain was localized by two-dimensional peptide mapping to a serine residue(s) in a single major tryptic phosphopeptide. This phosphopeptide co-migrated with a tryptic peptide that was produced following in vitro phosphorylation of platelet myosin heavy chain using protein kinase C. The sites phosphorylated in the 20-kDa myosin light chain in intact cells were analyzed by two-dimensional mapping of tryptic peptides and found to correspond to Ser1 and Ser2 in the turkey gizzard myosin light chain. In vitro phosphorylation of purified human platelet myosin by protein kinase C showed that in addition to Ser1 and Ser2, a third site corresponding to Thr9 in turkey gizzard myosin light chain is also phosphorylated. The phosphorylatable myosin light chains from human platelets were found to consist of two major isoforms present in approximately equal amounts, but differing in their molecular weights and isoelectric points. A third, minor isoform was also visualized by two-dimensional gel electrophoresis. Following treatment with TPA, both the mono- and diphosphorylated forms of each isoform could be visualized, and the sites of phosphorylation were identified. The phosphate content rose from negligible amounts found prior to treatment with TPA to 1.2 mol of phosphate/mol of myosin light chain and 0.7 mol of phosphate/mol of myosin heavy chain following treatment. These results suggest that TPA mediates phosphorylation of both myosin light and heavy chains in intact platelets by activation of protein kinase C.  相似文献   

9.
The distribution of nonmuscle myosin isoforms in brain and aorta was studied by using polyclonal antibodies against two synthetic peptides selected from a region near the carboxyl terminus of bovine brain (peptide IIB) and human macrophage (peptide IIA) myosin. Immunoblots of brain homogenates and purified myosin showed two major bands stained by anti-peptide IIB (MIIB1 and MIIB2) and a minor band stained by anti-peptide IIA (MIIA2). Polyclonal anti-human platelet myosin antibodies did not react with MIIB isoforms. In cryosections from bovine, rat, and mouse brains, anti-peptide IIB stained most neuronal cells. In bovine cryosections, glial staining was also observed. In contrast, anti-peptide IIA and anti-platelet myosin antibodies primarily stained blood vessels. In bovine aorta, the anti-peptide antibodies recognized four bands, MIIB3, MIIB4, MIIA1, and MIIA2. Only MIIA2 was recognized by anti-human platelet myosin antibodies. In bovine aorta cryosections, anti-peptide IIB stained smooth muscle cells in tunica intima and tunica media but did not stain endothelial cells. Anti-peptide IIA stained smooth muscle cells in the tunica media, and endothelial cells of vaso vasorum but not of aorta. Only polyclonal anti-platelet myosin antibodies stained the endothelial cells of aorta tunica intima. These results indicate that multiple isoforms of cellular myosins exist in mammals, that these isoforms are expressed in a cell specific manner, and that the major myosin isoforms isolated from whole brain originate from neurons and, at least in bovine brain, from glia, but not from blood vessels.  相似文献   

10.
Smooth muscle cells express isoforms of actin and myosin heavy chains (MHC). In early postnatal animals the nonmuscle (NM) actin and MHC isoforms in vascular (aorta) smooth muscle were present in relatively high percentages. More than 30% of the MHC and 40% of the actin isoforms were NM. The relative percentage of the NM isoforms decreased significantly as the animals reached maturity, with NM MHC less than 10% and NM actin less than 30% of the totals. Concurrent with this decrease in NM isoforms was an increase in the smooth muscle (SM) isoforms. The relative changes and time frame in which these changes occurred were very similar for the actin and MHC isoforms. In arterial tissue there were species differences for changes with development in the two SM MHC isoforms (SM1 and SM2). The ratio of SM1:SM2 in young rat aorta was approximately 0.5, while this same ratio was approximately 3 in young swine carotid. Both adult rats and swine had a SM1:SM2 MHC ratio of approximately 1.2. Rat bladder smooth muscle showed no significant change in NM vs SM ratio between young and old rats, while the SM1:SM2 ratio decreased from 2.7 to 1.7 between these age groups. The shifts in alpha and beta actin were similar to those in the vascular tissue, but of much smaller magnitude.  相似文献   

11.
Vascular smooth muscle caldesmon   总被引:10,自引:0,他引:10  
Caldesmon, a major actin- and calmodulin-binding protein, has been identified in diverse bovine tissues, including smooth and striated muscles and various nonmuscle tissues, by denaturing polyacrylamide gel electrophoresis of tissue homogenates and immunoblotting using rabbit anti-chicken gizzard caldesmon. Caldesmon was purified from vascular smooth muscle (bovine aorta) by heat treatment of a tissue homogenate, ion-exchange chromatography, and affinity chromatography on a column of immobilized calmodulin. The isolated protein shared many properties in common with chicken gizzard caldesmon: immunological cross-reactivity, Ca2+-dependent interaction with calmodulin, Ca2+-independent interaction with F-actin, competition between actin and calmodulin for caldesmon binding only in the presence of Ca2+, and inhibition of the actin-activated Mg2+-ATPase activity of smooth muscle myosin without affecting the phosphorylation state of myosin. Maximal binding of aorta caldesmon to actin occurred at 1 mol of caldesmon: 9-10 mol of actin, and binding was unaffected by tropomyosin. Half-maximal inhibition of the actin-activated myosin Mg2+-ATPase occurred at approximately 1 mol of caldesmon: 12 mol of actin. This inhibition was also unaffected by tropomyosin. Caldesmon had no effect on the Mg2+-ATPase activity of smooth muscle myosin in the absence of actin. Bovine aorta and chicken gizzard caldesmons differed in several respects: Mr (149,000 for bovine aorta caldesmon and 141,000 for chicken gizzard caldesmon), extinction coefficient (E1%280nm = 19.5 and 5.0 for bovine aorta and chicken gizzard caldesmon, respectively), amino acid composition, and one-dimensional peptide maps obtained by limited chymotryptic and Staphylococcus aureus V8 protease digestion. In a competitive enzyme-linked immunosorbent assay, using anti-chicken gizzard caldesmon, a 174-fold molar excess of bovine aorta caldesmon relative to chicken gizzard caldesmon was required for half-maximal inhibition. These studies establish the widespread tissue and species distribution of caldesmon and indicate that vascular smooth muscle caldesmon exhibits physicochemical differences yet structural and functional similarities to caldesmon isolated from chicken gizzard.  相似文献   

12.
Nonmuscle and smooth muscle myosin isoforms in bovine endothelial cells   总被引:3,自引:0,他引:3  
A panel of monoclonal antibodies, specific for human platelet (NM-A9, NM-F6, and NM-G2) and for bovine smooth muscle (SM-E7) myosin heavy chains (MHC), were used to study the composition and the distribution of myosin isoforms in bovine endothelial cells (EC), in vivo and in vitro. Using indirect and double immunofluorescence techniques, we have found that in the intact aortic endothelium there is expression of nonmuscle MHC (NM-MHC), exclusively. By contrast, hepatic sinusoidal endothelium as well as cultured bovine aortic EC (BAEC) in the subconfluent phase of growth show coexistence of NM- and smooth muscle MHC (SM-MHC) isoforms. SM myosin immunoreactivity disappears when cultured BAEC become confluent. In this phase of cell growth, NM-MHC isoforms are localized differently within the cells, i.e., in the cytoplasm around the nucleus or in the cortical, submembranous region of EC cytoplasm. A third type of intracellular distribution of NM-MHC immunoreactivity was evident in the cell periphery of binucleated, confluent BAEC. These data indicate that (1) several myosin isoforms are differently distributed in bovine endothelia; and (2) SM myosin expression and the specific subcellular localization of NM myosin isoforms within EC might be regulated by cell-cell interactions.  相似文献   

13.
Two distinct cDNA clones for nonmuscle myosin heavy chain (MHC) were isolated from a chicken fibroblast cDNA library by cross-hydridization under a moderate stringency with chicken gizzard smooth muscle MHC cDNA. These two fibroblast MHC and the gizzard MHC are each encoded in different genes in the chicken genome. Northern blot analysis showed that both of the nonmuscle MHC mRNAs were expressed not only in fibroblasts but also in a variety of tissues including brain, lung, kidney, spleen, and skeletal, cardiac and smooth muscles. However, the relative contents of the two nonmuscle MHC mRNAs varied greatly among tissues. The encoded amino acid sequences of the nonmuscle MHCs were highly similar to each other (81% identity) and to the smooth muscle MHC (81-84%), but much less similar to vertebrate skeletal muscle MHCs (38-41%) or to protista nonmuscle MHCs (35-36%). A phylogenic tree of MHC isoforms was constructed by calculating the similarity scores between these MHC sequences. An examination of the tree showed that the vertebrate sarcomeric (skeletal and cardiac) MHC isoforms are encoded in a very closely related multigene family, and that the vertebrate non-sarcomeric (smooth muscle and nonmuscle) MHC isoforms define a distinct, less conserved MHC gene family.  相似文献   

14.
Nonmuscle myosin can generate force and shortening in smooth muscle, as revealed by studies of the urinary bladder from mice lacking smooth muscle myosin heavy chain (SM-MHC) but expressing the nonmuscle myosin heavy chains A and B (NM-MHC A and B; Morano, I., G.X. Chai, L.G. Baltas, V. Lamounier-Zepter, G. Lutsch, M. Kott, H. Haase, and M. Bader. 2000. Nat. Cell Biol. 2:371-375). Intracellular calcium was measured in urinary bladders from SM-MHC-deficient and SM-MHC-expressing mice in relaxed and contracted states. Similar intracellular [Ca2+] transients were observed in the two types of preparations, although the contraction of SM-MHC-deficient bladders was slow and lacked an initial peak in force. The difference in contraction kinetics thus do not reflect differences in calcium handling. Thick filaments were identified with electron microscopy in smooth muscle cells of SM-MHC-deficient bladders, showing that NM-MHC can form filaments in smooth muscle cells. Maximal shortening velocity of maximally activated, skinned smooth muscle preparations from SM-MHC-deficient mice was significantly lower and more sensitive to increased MgADP compared with velocity of SM-MHC-expressing preparations. Active force was significantly lower and less inhibited by increased inorganic phosphate. In conclusion, large differences in nucleotide and phosphate binding exist between smooth and nonmuscle myosins. High ADP binding and low phosphate dependence of nonmuscle myosin would influence both velocity of actin translocation and force generation to promote slow motility and economical force maintenance of the cell.  相似文献   

15.
The actin-regulatory protein profilin has been shown to regulate the actin cytoskeleton and the motility of nonmuscle cells. To test the hypothesis that profilin plays a role in regulating smooth muscle contraction, profilin antisense or sense oligodeoxynucleotides were introduced into the canine carotid smooth muscle by a method of reversible permeabilization, and these strips were incubated for 2 days for protein downregulation. The treatment of smooth muscle strips with profilin antisense oligodeoxynucleotides inhibited the expression of profilin; it did not influence the expression of actin, myosin heavy chain, and metavinculin/vinculin. Profilin sense did not affect the expression of these proteins in smooth muscle tissues. Force generation in response to stimulation with norepinephrine or KCl was significantly lower in profilin antisense-treated muscle strips than in profilin sense-treated strips or in muscle strips not treated with oligodeoxynucleotides. The depletion of profilin did not attenuate increases in phosphorylation of the 20-kDa regulatory light chain of myosin (MLC20) in response to stimulation with norepinephrine or KCl. The increase in F-actin/G-actin ratio during contractile stimulation was significantly inhibited in profilin-deficient smooth muscle strips. These results suggest that profilin is a necessary molecule of signaling cascades that regulate carotid smooth muscle contraction, but that it does not modulate MLC20 phosphorylation during contractile stimulation. Profilin may play a role in the regulation of actin polymerization or organization in response to contractile stimulation of smooth muscle.  相似文献   

16.
It has been reported that catch and striated muscle myosin heavy chains of scallop are generated through alternative splicing from a single gene [Nyitray et al. (1994) Proc. Natl. Acad. Sci. USA 91, 12686-12690]. They suggested that the catch muscle type myosin was expressed in various tissues of scallop, including the gonad, heart, foot, and mantle. However, there have been no reports of the primary structure of myosin from tissues other than the adductor muscles. In this study, we isolated a cDNA encoding the motor domain of myosin from the mantle tissue of scallop (Patinopecten yessoensis), and determined its nucleotide sequence. Sequence analysis revealed that mantle myosin exhibited 65% identity with Drosophila non muscle myosin, 60% with chicken gizzard smooth muscle myosin, and 44% with scallop striated muscle myosin. The mantle myosin has inserted sequences in the 27 kDa domain of the head region, and has a longer loop 1 structure than those of scallop striated and catch muscle myosins. Phylogenetic analysis suggested that the mantle myosin is classified as a smooth/nonmuscle type myosin. Western blot analysis with antibodies produced against the N-terminal region of the mantle myosin revealed that this myosin was specifically expressed in the mantle pallial cell layer consisting of nonmuscle cells. Our results show that mantle myosin is classified as a nonmuscle type myosin in scallop.  相似文献   

17.
Rho family small GTPases (Rho, Rac, and Cdc42) play an important role in cell motility, adhesion, and cell division by signaling reorganization of the actin cytoskeleton. Here, we report an isoactin-specific, Rho GTPase-dependent signaling cascade in cells simultaneously expressing smooth muscle and nonmuscle actin isoforms. We transfected primary cultures of microvascular pericytes, cells related to vascular smooth muscle cells, with various Rho-related and Rho-specific expression plasmids. Overexpression of dominant positive Rho resulted in the formation of nonmuscle actin-containing stress fibers. At the same time, -vascular smooth muscle actin (VSMactin) containing stress fibers were disassembled, resulting in a dramatic reduction in cell size. Rho activation also yielded a disassembly of smooth muscle myosin and nonmuscle myosin from stress fibers. Overexpression of wild-type Rho had similar but less dramatic effects. In contrast, dominant negative Rho and C3 exotransferase or dominant positive Rac and Cdc42 expression failed to alter the actin cytoskeleton in an isoform-specific manner. The loss of smooth muscle contractile protein isoforms in pericyte stress fibers, together with a concomitant decrease in cell size, suggests that Rho activation influences "contractile" phenotype in an isoactin-specific manner. This, in turn, should yield significant alteration in microvascular remodeling during developmental and pathologic angiogenesis. vascular smooth muscle actin; Rho GTPase; pericyte; myosin; cytoskeleton  相似文献   

18.
The mechanism of the ATPase [EC 3.6.1.3] reaction of porcine platelet myosin and the binding properties of platelet myosin with rabbit skeletal muscle F-actin were investigated. The kinetic properties of the platelet myosin ATPase reaction, that is, the rate, the extent of fluorescence enhancement of myosin, the size of the initial P1 burst of myosin, and the amount of nucleotides bound to myosin during the ATPase reaction, were very similar to those found for other myosins. Strong binding of platelet myosin with rabbit skeletal muscle F-actin, as found for smooth muscle myosin, was suggested by the following results. The rate of the ATP-induced dissociation of hybrid actomyosin, reconstituted from platelet myosin and skeletal muscle F-actin, was very slow. The amount of ATP necessary for complete dissociation of hybrid actomyosin was 2 mol/mol of myosin, although skeletal muscle actomyosin is known to dissociate completely upon addition of 1 mol ATP per mol of myosin. Unlike skeletal muscle myosin, the EDTA(K+)-ATPase activity of platelet myosin was inhibited by skeletal muscle F-actin. These observations indicate that ATP hydrolysis by vertebrate nonmuscle myosin follows the same mechanism as with other myosins and that the binding properties of nonmuscle myosin with F-actin are similar to those of smooth muscle myosin but not to those of skeletal muscle myosin.  相似文献   

19.
20.
Two-dimensional mapping of the tryptic phosphopeptides generated following in vitro protein kinase C phosphorylation of the myosin heavy chain isolated from human platelets and chicken intestinal epithelial cells shows a single radioactive peptide. These peptides were found to comigrate, suggesting that they were identical, and amino acid sequence analysis of the human platelet tryptic peptide yielded the sequence -Glu-Val-Ser-Ser(PO4)-Leu-Lys-. Inspection of the amino acid sequence for the chicken intestinal epithelial cell myosin heavy chain (196 kDa) derived from cDNA cloning showed that this peptide was identical with a tryptic peptide present near the carboxyl terminal of the predicted alpha-helix of the myosin rod. Although other vertebrate nonmuscle myosin heavy chains retain neighboring amino acid sequences as well as the serine residue phosphorylated by protein kinase C, this residue is notably absent in all vertebrate smooth muscle myosin heavy chains (both 204 and 200 kDa) sequenced to date.  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