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.
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   

2.
Isolation of a non-muscle myosin heavy chain gene from Acanthamoeba   总被引:6,自引:0,他引:6  
We have isolated a non-muscle myosin heavy chain gene from Acanthamoeba castellanii using as a heterologous probe a sarcomeric myosin heavy chain gene from Caenorhabditis elegans. The amoeba genomic clone has been tentatively identified as containing a myosin II heavy chain gene based on hybridization to a 5300-nucleotide RNA species, hybrid selection of a mRNA encoding a 185-kDa polypeptide, specific immunoprecipitation of this polypeptide with antiserum to myosin II, and an exact match between the DNA sequence and a carboxyl-terminal myosin II peptide previously sequenced by protein chemical methods (C?té, G.P., Robinson, E.A., Appella, E., and Korn, E. D. (1984) J. Biol. Chem. 259, 12781-12787). We also sequenced a region of the gene whose deduced amino acid sequence shows strong homology with that region of muscle myosins which is thought to be involved in nucleotide binding. These results indicate that the amoeba genomic clone contains at least 90% of the coding information for the 185-kDa heavy chain polypeptide and that the bulk of the gene contains very little intron DNA. Genomic blots of amoeba DNA probed with a portion of this myosin gene indicate the presence of additional highly related sequences within the amoeba genome.  相似文献   

3.
We have completely sequenced a gene encoding the heavy chain of myosin II, a nonmuscle myosin from the soil ameba Acanthamoeba castellanii. The gene spans 6 kb, is split by three small introns, and encodes a 1,509-residue heavy chain polypeptide. The positions of the three introns are largely conserved relative to characterized vertebrate and invertebrate muscle myosin genes. The deduced myosin II globular head amino acid sequence shows a high degree of similarity with the globular head sequences of the rat embryonic skeletal muscle and nematode unc 54 muscle myosins. By contrast, there is no unique way to align the deduced myosin II rod amino acid sequence with the rod sequence of these muscle myosins. Nevertheless, the periodicities of hydrophobic and charged residues in the myosin II rod sequence, which dictate the coiled-coil structure of the rod and its associations within the myosin filament, are very similar to those of the muscle myosins. We conclude that this ameba nonmuscle myosin shares with the muscle myosins of vertebrates and invertebrates an ancestral heavy chain gene. The low level of direct sequence similarity between the rod sequences of myosin II and muscle myosins probably reflects a general tolerance for residue changes in the rod domain (as long as the periodicities of hydrophobic and charged residues are largely maintained), the relative evolutionary "ages" of these myosins, and specific differences between the filament properties of myosin II and muscle myosins. Finally, sequence analysis and electron microscopy reveal the presence within the myosin II rodlike tail of a well-defined hinge region where sharp bending can occur. We speculate that this hinge may play a key role in mediating the effect of heavy chain phosphorylation on enzymatic activity.  相似文献   

4.
We have previously purified and characterized a Dictyostelium myosin II heavy chain kinase which phosphorylates threonine residues (C?té, G. P., and Bukiejko, U. (1987) J. Biol. Chem. 262, 1065-1072). The phosphorylated threonines are located within a 34-kDa fragment which can be selectively cleaved from the carboxyl terminal end of the Dictyostelium myosin II tail. Tryptic and chymotryptic digests of the 34-kDa fragment phosphorylated with the kinase have now been performed and the resulting phosphopeptides isolated and sequenced. Two phosphorylated threonine residues have been identified, corresponding to residues 1833 and 2029 in the complete amino acid sequence of the Dictyostelium myosin II heavy chain. These amino acids are 87 and 283 residues, respectively, distant from the carboxyl terminus of the Dictyostelium myosin II heavy chain and are present in sections of the tail which seem to be alpha-helical coiled coils. In contrast, the three Acanthamoeba myosin II heavy chain phosphorylation sites are located within 10 residues of each other in a small globular domain at the carboxyl terminal tip of the tail (C?té, G. P., Robinson, E. A., Appella, E., and Korn, E. D. (1984) J. Biol. Chem. 259, 12781-12787). This suggests that the mechanism by which heavy chain phosphorylation inhibits the actin-activated ATPase activity and filament-forming properties of the two myosins may be quite different.  相似文献   

5.
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   

6.
Sarcomeric myosin heavy chain (MyHC) is the major contractile protein of striated muscle. Six tandemly linked skeletal MyHC genes on chromosome 17 and two cardiac MyHC genes on chromosome 14 have been previously described in the human genome. We report the identification of three novel human sarcomeric MyHC genes on chromosomes 3, 7, and 20, which are notable for their atypical size and intron-exon structure. Two of the encoded proteins are structurally most like the slow-beta MyHC, whereas the third one is closest to the adult fast IIb isoform. Data from pairwise comparisons of aligned coding sequences imply the existence of ancestral genomes with four sarcomeric genes before the emergence of a dedicated smooth muscle MyHC gene. To further address the evolutionary relationships of the distinct sarcomeric and nonsarcomeric rod sequences, we have identified and further annotated human genomic DNA sequences corresponding to 14 class-II MyHCs. An extensive analysis provides a timeline for intron gain and loss, gene contraction and expansion, and gene conversion among genes encoding class-II myosins. One of the novel human genes is found to have introns at positions shared only with the molluscan catchin/MyHC gene, providing evidence for the structure of a pre-Cambrian ancestral gene.  相似文献   

7.
The heavy chain fragments generated by restricted proteolysis of the smooth chicken gizzard myosin subfragment-1 (S-1) with trypsin, Staphylococcus aureus V8 protease, and chymotrypsin were isolated and submitted to partial amino acid sequencing. The comparison between the smooth and striated muscle myosin sequences permitted the unambiguous structural characterization of the two protease-vulnerable segments joining the three putative domain-like regions of the smooth head heavy chain. The smooth carboxyl-terminal connector is a serine-rich region located around positions 632-640 of the rabbit skeletal sequence and would represent the "A" site that is conformationally sensitive to the myosin 10 S-6 transition and to its interaction with actin (Ikebe, M., and Hartshorne, D. J. (1986) Biochemistry 25, 6177-6185). A third site which undergoes a nucleotide-dependent chymotryptic cleavage which inactivates the Mg2+-ATPase (Okamoto, Y., and Sekine, T. (1981) J. Biochem. (Tokyo) 90, 833-842, 843-849) was identified at Trp-31/Ser-32. It is vicinal to Lys-34 that is monomethylated in the skeletal heavy chain but not at all in the smooth sequence. However, the two trimethyl lysine residues present in the skeletal sequence are conserved in the same regions of the smooth S-1 and may play a general functional role in myosin. The smooth central 50-kDa segment could be selectively destroyed by a mild tryptic digestion in the absence of any unfolding agent, with a concomitant inhibition of the ATPase activities. This feature is in line with the proposed domain structure of the S-1 heavy chain and also suggests a relationship between the specific biochemical properties of the smooth S-1 and the particular conformation of its 50-kDa region.  相似文献   

8.
The two cardiac myosin heavy chain isoforms, alpha and beta, differ functionally, alpha Myosin exhibits higher actin-activated ATPase than does beta myosin, and hearts expressing alpha myosin exhibit increased contractility relative to hearts expressing beta myosin. To understand the molecular basis for this functional difference, we determined the complete nucleotide sequence of full-length rat alpha and beta myosin heavy chain cDNAs. This study represents the first opportunity to compare full-length fast ATPase and slow ATPase muscle myosin sequences. The alpha and beta myosin heavy chain amino acid sequences are more related to each other than to other sarcomeric myosin heavy chain sequences. Of the 1938 amino acid residues in alpha and beta myosin heavy chain, 131 are non-identical with 37 non-conservative changes. Two-thirds of these non-identical residues are clustered, and several of these clusters map to regions that have been implicated as functionally important. Some of the regions identified by the clusters of non-identical amino acid residues may affect actin binding, ATP hydrolysis and force production.  相似文献   

9.
Vertebrate smooth muscle myosin heavy chains (MHCs) exist as two isoforms with molecular masses of 204 and 200 kDa (MHC204 and MHC200) that are generated from a single gene by alternative splicing of mRNA (Nagai, R., Kuro-o, M., Babij, P., and Periasamy, M. (1989) J. Biol. Chem. 264, 9734-9737). A dimer of two MHCs associated with two pairs of myosin light chains forms a functional myosin molecule. To investigate the isoform composition of the MHCs in native myosin, antibodies specific for MHC204 were generated and used to immunoprecipitate purified bovine aortic smooth muscle myosin from a solution containing equal amounts of each isoform. MHC204 quantitatively removed from this mixture was completely free of MHC200. Immunoprecipitation of the supernatant with an antiserum that recognizes both isoforms equally well revealed that only MHC200 remained. We conclude that only homodimers of MHC204 and MHC200 exist under these conditions. A method is described for the purification of enzymatically active MHC204 and myosin on a protein G-agarose high performance liquid chromatography column containing immobilized MHC204 antibodies. We show, using an in vitro motility assay, that the movement of actin filaments by myosin containing 204-kDa heavy chains (0.435 +/- 0.115 microns/s) was not significantly different from that of myosin containing 200-kDa heavy chains (0.361 +/- 0.078 microns/s) or from myosin containing equal amounts of each heavy chain isoform (0.347 +/- 0.082 microns/s).  相似文献   

10.
The amino acid sequence of the 50-kDa fragment that is released by limited tryptic digestion of the head portion of rabbit skeletal muscle myosin was determined by analysis and alignment of sets of peptides generated by digestion of the fragment at arginine or methionine residues. This fragment contains residues 205-636 of the myosin heavy chain; among the residues of particular interest in this fragment are N epsilon-trimethyllysine, one of four methyl-amino acids in myosin, and Ser-324, which is photoaffinity labeled by an ATP analogue (Mahmood, R., Elzinga, M., and Yount, R. G. (1989) Biochemistry 28, 3989-3995). Combination of this sequence with those of the 23- and 20-kDa fragments yields an 809-residue sequence that constitutes most of the heavy chain of chymotryptic S-1 of this myosin.  相似文献   

11.
The nucleotide sequence of the cDNA encoding myosin heavy chain of chum salmon Oncorhynchus keta fast skeletal muscle was determined. The sequence consists of 5,994 bp, including 5,814 bp of translated region deducing an amino acid sequence of 1,937 residues. The deduced sequence showed 79% homology to that of rabbit fast skeletal myosin and 84-87% homology to those of fast skeletal myosins from walleye pollack, white croaker and carp. The putative binding-sites for ATP, actin and regulatory light-chains in the subfragment-1 region of the salmon myosin showed high homology with the fish myosins (78-100% homology). However, the Loop-1 and Loop-2 showed considerably low homology (31-60%). On the other hand, the deduced sequences of subfragment-2 (533 residues) and light meromyosin (564 residues) showed 88-93% homology to the corresponding regions of the fish myosins. It becomes obvious that several specific residues of the rabbit LMM are substituted to Gly in the salmon LMM as well as the other fish LMMs. This may be involved in the structural instability of the fish myosin tail region.  相似文献   

12.
To probe the molecular properties of the actin recognition site on the smooth muscle myosin heavy chain, the rigor complexes between skeletal F-actin and chicken gizzard myosin subfragments 1 (S1) were investigated by limited proteolysis and by chemical cross-linking with 1-ethyl-3-[3-(dimethyl-amino)propyl]carbodiimide. Earlier, these approaches were used to analyze the actin site on the skeletal muscle myosin heads [Mornet, D., Bertrand, R., Pantel, P., Audemard, E., & Kassab, R. (1981) Biochemistry 20, 2110-2120; Labbé, J.P., Mornet, D., Roseau, G., & Kassab, R. (1982) Biochemistry 21, 6897-6902]. In contrast to the case of the skeletal S1, the cleavage with trypsin or papain of the sensitive COOH-terminal 50K-26K junction of the head heavy chain had no effect on the actin-stimulated Mg2+-ATPase activity of the smooth S1. Moreover, actin binding had no significant influence on the proteolysis at this site whereas it abolished the scission of the skeletal S1 heavy chain. The COOH-terminal 26K segment of the smooth papain S1 heavy chain was converted by trypsin into a 25K peptide derivative, but it remained intact in the actin-S1 complex. A single actin monomer was cross-linked with the carbodiimide reagent to the intact 97K heavy chain of the smooth papain S1. Experiments performed on the complexes between F-actin and the fragmented S1 indicated that the site of cross-linking resides within the COOH-terminal 25K fragment of the S1 heavy chain. Thus, for both the striated and smooth muscle myosins, this region appears to be in contact with F-actin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The active site of chicken gizzard myosin was labeled by direct photoaffinity labeling with [3H]UDP. [3H] UDP was stably trapped at the active site by addition of vanadate (Vi) and Co2+. The extraordinary stability of the myosin.Co2+.[3H]UDP.Vi complex (t1/2 greater than 5 days at 0 degrees C) allowed it to be purified free of extraneous [3H]UDP before irradiation began. Upon UV irradiation, greater than 60% of the trapped [3H]UDP was photoincorporated into the active site. Only the 200-kDa heavy chain was labeled, confirming earlier results (Maruta, H., and Korn, E. (1981) J. Biol. Chem. 256, 499-502) using [3H]UTP. Extensive tryptic digestion of photolabeled myosin subfragment 1 followed by high performance liquid chromatography separations and removal of nucleotide phosphates by treatment with alkaline phosphatase allowed two labeled peptides to be isolated. Sequencing of the labeled peptides and radioactive counting showed that Glu185 was the residue labeled. Since UDP is a "zero-length" cross-linker, Glu185 is located at the purine-binding pocket of the active site of smooth myosin and adjacent to the glycine-rich loop which binds the polyphosphate portion of ATP. This Glu residue is conserved in smooth and nonmuscle myosins and is the same residue identified previously by [3H]UTP photolabeling in Acanthamoeba myosin II (Atkinson, M. A., Robinson, E. A., Appella, E., and Korn, E. D. (1986) J. Biol. Chem. 261, 1844-1848).  相似文献   

14.
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   

15.
The N-terminal region of myosin's rod-like subfragment 2 (S2) joins the two heads of this dimeric molecule and is key to its function. Previously, a crystal structure of this predominantly coiled-coil region was determined for a short fragment (51 residues plus a leucine zipper) of the scallop striated muscle myosin isoform. In that study, the N-terminal 10-14 residues were found to be disordered. We have now determined the structure of the same scallop peptide in three additional crystal environments. In each of two of these structures, improved order has allowed visualization of the entire N-terminus in one chain of the dimeric peptide. We have also compared the melting temperatures of this scallop S2 peptide with those of analogous peptides from three other isoforms. Taken together, these experiments, along with examination of sequences, point to a diminished stability of the N-terminal region of S2 in regulated myosins, compared with those myosins whose regulation is thin filament linked. It seems plain that this isoform-specific instability promotes the off-state conformation of the heads in regulated myosins. We also discuss how myosin isoforms with varied thermal stabilities share the basic capacity to transmit force efficiently in order to produce contraction in their on states.  相似文献   

16.
In the preceding paper [Maita, T., Miyanishi, T., Matsuzono, K., Tanioka, Y., & Matsuda, G. (1991) J. Biochem. 110, 68-74], we reported the amino-terminal 837-residue sequence of the heavy chain of adult chicken pectoralis muscle myosin. This paper describes the carboxyl terminal 1,097-residue sequence and the linkage of the two sequences. Rod obtained by digesting myosin filaments with alpha-chymotrypsin was redigested with the protease at high KCl concentration, and two fragments, subfragment-2 and light meromyosin, were isolated and sequenced by conventional methods. The linkage of the two fragments was deduced from the sequence of an overlapping peptide obtained by cleaving the rod with cyanogen bromide. The rod contained 1,039 amino acid residues, but lacked the carboxyl-terminal 58 residues of the heavy chain. A carboxyl-terminal 63-residue peptide obtained by cleaving the whole heavy chain with cyanogen bromide was sequenced. Thus, the carboxyl terminal 1,097-residue sequence of the heavy chain was completed. The linkage of subfragment-1 and the rod was deduced from the sequence of an overlapping peptide between the two which was obtained by cleaving heavy meromyosin with cyanogen bromide. Comparing the sequence of the adult myosin thus determined with that of chicken embryonic myosin reported by Molina et al. [Molina, M.I., Kropp, K.E., Gulick, J., & Robbins, J. (1987) J. Biol. Chem. 262, 6478-6488], we found that the sequence homology is 94%.  相似文献   

17.
Myosins are represented by a wide range of different classes of molecule, of which the most extensively studied are the class II myosins which drive muscle contraction and cell organization; the functional unit of class II myosins comprises two myosin heavy chains (MHCs). This minireview gives an update on class II MHCs of nematodes and describes a comparative analysis of MHC genes from nematodes and other organismal groups. Genetic analyses of sequence data for the four functional domains of MHCs (i.e., the SH3-like N-terminal, head, neck and tail domains) reveal a delineation between both the nematode and non-nematode myosins and between muscle and non-muscle myosins. The distinctiveness of the MHCs of nematodes suggests functional and tissue specialization. The elucidation of the functional roles of myosins and other molecules in specific signaling pathways in nematodes has the potential to lead to new intervention strategies for parasites via the specific disruption or interruption of key developmental processes, having biotechnological implications in the longer term.  相似文献   

18.
Protein kinase C phosphorylated both the 19/21-kDa regulatory light chains and heavy chains of bovine brain myosin. The major phosphorylation sites of the light chains were on their threonyl residues, while those for myosin light chain kinase were on their seryl residues. Whereas several non-muscle regular myosins have been reported to be phosphorylated by different types of protein kinases at the non-helical small segments at the tail ends of the heavy chains, the phosphorylation sites for protein kinase C were localized on the head portion of the heavy chains of brain myosin. The possible role of phosphorylation of brain myosin by protein kinase C in the regulation of motility of neural cells is discussed.  相似文献   

19.
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   

20.
Su Z  Kiehart DP 《Biochemistry》2001,40(12):3606-3614
Conventional myosins (myosin-IIs) generate forces for cell shape change and cell motility. Myosin heavy chain phosphorylation regulates myosin function in simple eukaryotes and may also be important in metazoans. To investigate this regulation in a complex eukaryote, we purified the Drosophila myosin-II tail expressed in Escherichia coli and showed that it was phosphorylated in vitro by protein kinase C(PKC) at serines 1936 and 1944, which are located in the nonhelical globular tail piece. These sites are close to a conserved serine that is phosphorylated in vertebrate, nonmuscle myosin-IIs. If the two serines are mutagenized to alanine or aspartic acid, phosphorylation no longer occurs. Using a 341 amino acid tail fragment, we show that there is no difference in the salt-dependent assembly of wild-type phosphorylated and mutagenized polypeptides. Thus, the nonmuscle myosin heavy chain in Drosophila, which is encoded by the zipper gene, appears to be similar to rabbit nonmuscle myosin-IIA. In vivo, we generated transgenic flies that expressed the various myosin heavy chain variants in a zipper null or near-null genetic background. Like their wild-type counterparts, such variants are able to completely rescue the lethal phenotype due to severe zipper mutations. These results suggest that while the myosin-II heavy chain can be phosphorylated by PKC, regulation by this enzyme is not required for viability in Drosophila. Conservation during 530-1000 million years of evolution suggests that regulation by heavy chain phosphorylation may contribute to nonmuscle myosin-II function in some real, but minor, way.  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