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.
Several actin-binding proteins participate in the morphological changes that occur during amoeboid movement. The gene encoding one of these proteins, the gelation factor ABP-120, was identified and characterized from trophozoites of Entamoeba histolytica . The sequence contains 2574 nucleotides, with an open reading frame of 858 amino acids, giving a protein of 93 kDa belonging to the spectrin family. The N-terminal domain of ABP-120 from E. histolytica revealed a consensus site for actin binding homologous to the actin-binding sites of ABP-120 of Dictyostelium discoideum , α-actinin and spectrin. Analysis of the central domain revealed the presence of four repeats of a 73-amino-acid motif constituting 31% of the protein. In addition, a stretch of 105 amino acids was highly divergent when compared with the C-terminal domain of D. discoideum ABP-120. This sequence showed short motifs that are homologous to microtubule-binding domains. We found that ABP-120 from E. histolytica binds to F-actin. In addition, upon motility of the parasite, this protein localized in the pseudopod and the uroid region, implying a role for ABP-120 in movement and capping of surface receptors in E. histolytica .  相似文献   

2.
The 120-kD gelation factor and alpha-actinin are among the most abundant F-actin cross-linking proteins in Dictyostelium discoideum. Both molecules are homodimers and have extended rod-like configurations that are respectively approximately 35 and 40 nm long. Here we report the complete cDNA sequence of the 120-kD gelation factor which codes for a protein of 857 amino acids. Its calculated molecular mass is 92.2 kD which is considerably smaller than suggested by its mobility in SDS-PAGE. Analysis of the sequence shows a region that is highly homologous to D. discoideum alpha-actinin, chicken fibroblast alpha-actinin, and human dystrophin. This conserved domain probably represents an actin binding site that is connected to the rod-forming part of the molecule via a highly charged stretch of amino acids. Whereas the sequence of alpha-actinin (Noegel, A., W. Witke, and M. Schleicher. 1987. FEBS [Fed. Eur. Biochem. Soc.] Lett. 221:391-396) suggests that the extended rod domain of the molecule is based on four spectrin-like repeats with high alpha-helix potential, the rod domain of the 120-kD gelation factor is constructed from six 100-residue repeats that have a high content of glycine and proline residues and which, in contrast to alpha-actinin, do not appear to have a high alpha-helical content. These repeats show a distinctive pattern of regions that have high beta-sheet potential alternating with short zones rich in residues with a high potential for turns. This observation suggests that each 100-residue motif has a cross-beta conformation with approximately nine sheets arranged perpendicular to the long axis of the molecule. In the high beta-potential zones every second residue is often hydrophobic. In a cross-beta structure, this pattern would result in one side of the domain having a surface rich in hydrophobic side chains which could account for the dimerization of the 120-kD gelation factor subunits.  相似文献   

3.
We have studied the effect of the Dictyostelium discoideum 30,000-D actin-bundling protein on the assembly and disassembly of pyrenyl-labeled actin in vitro. The results indicate that the protein is a potent inhibitor of the rate of actin depolymerization. The inhibition is rapid, dose dependent, and is observed at both ends of the filament. There is little effect of 30-kD protein on the initial rate of elongation from F-actin seeds or on the spontaneous nucleation of actin polymerization. We could detect little or no effect on the critical concentration. The novel feature of these results is that the filament ends are free for assembly but are significantly impaired in disassembly with little change in the critical concentration at steady state. The effects appear to be largely independent of the cross-linking of actin filaments by the 30-kD protein. Actin cross-linking proteins may not only cross-link actin filaments, but may also differentially protect filaments in cells from disassembly and promote the formation of localized filament arrays with enhanced stability.  相似文献   

4.
W Witke  M Schleicher  A A Noegel 《Cell》1992,68(1):53-62
We generated by gene disruption Dictyostelium cells that lacked both the F-actin cross-linking proteins, alpha-actinin and gelation factor. Several major cell functions, such as growth, chemotaxis, phagocytosis, and pinocytosis, were apparently unaltered. However, in all double mutants, development was greatly impaired. After formation of aggregates, cells were very rarely able to form fruiting bodies. This ability was rescued when mutant and wild-type strains were mixed in a ratio of 70 to 30. The developmental program in the mutant was not arrested, since the expression pattern of early and late genes remained unchanged. Development of the mutant was rendered normal when a functional alpha-actinin gene was introduced and expressed, showing the morphogenetic defect to be due to the absence of the two F-actin cross-linking proteins. These findings suggest the existence of a functional network allowing mutual complementation of certain actin-binding proteins.  相似文献   

5.
Dictyostelium discoideum plasma membranes isolated by each of three procedures bind F-actin. The interactions between these membranes and actin are examined by a novel application of falling ball viscometry. Treating the membranes as multivalent actin-binding particles analogous to divalent actin-gelation factors, we observe large increases in viscosity (actin cross-linking) when membranes of depleted actin and myosin are incubated with rabbit skeletal muscle F-actin. Pre- extraction of peripheral membrane proteins with chaotropes or the inclusion of Triton X-100 during the assay does not appreciably diminish this actin cross-linking activity. Lipid vesicles, heat- denatured membranes, proteolyzed membranes, or membranes containing endogenous actin show minimal actin cross-linking activity. Heat- denatured, but not proteolyzed, membranes regain activity when assayed in the presence of Triton X-100. Thus, integral membrane proteins appear to be responsible for some or all of the actin cross-linking activity of D. discoideum membranes. In the absence of MgATP, Triton X- 100 extraction of isolated D. discoideum membranes results in a Triton- insoluble residue composed of actin, myosin, and associated membrane proteins. The inclusion of MgATP before and during Triton extraction greatly diminishes the amount of protein in the Triton-insoluble residue without appreciably altering its composition. Our results suggest the existence of a protein complex stabilized by actin and/or myosin (membrane cytoskeleton) associated with the D. discoideum plasma membrane.  相似文献   

6.
Reflection interference contrast microscopy combined with digital image processing was applied to study the motion of Dictyostelium discoideum cells in their pre-aggregative state on substrata of different adhesiveness (glass, albumin-covered glass, and freshly cleaved mica). The temporal variations of the size and shape of the cell/substratum contact area and the time course of advancement of pseudopods protruding in contact with the substratum were analyzed. The major goal was to study differences between the locomotion of wild-type cells and strains of triple mutants deficient in two F-actin cross-linking proteins (alpha-actinin and the 120-kDa gelation factor) and one F-actin fragmenting protein (severin). The size of contact area, AC, of both wild-type and mutant cells fluctuates between minimum and maximum values on the order of minutes, pointing toward an intrinsic switching mechanism associated with the mechanochemical control system. The fluctuation amplitudes are much larger on freshly cleaved mica than on glass. Wild-type and mutant cells exhibit remarkable differences on mica but not on glass. These differences comprise the population median of AC and alterations in pseudopod protrusion. AC is smaller by a factor of two or more for all mutants. Pseudopods protrude slower and shorter in the mutants. It is concluded that cell shape and pseudopods are destabilized by defects in the actin-skeleton, which can be overcompensated by strongly adhesive substrata. Several features of amoeboid cell locomotion on substrata can be understood on the basis of the minimum bending energy concept of soft adhering shells and by assuming that adhesion induces local alterations of the composite membrane consisting of the protein/lipid bilayer on the cell surface and the underlying actin-cortex.  相似文献   

7.
To determine the specific contribution of cytoskeletal proteins to cellular viscoelasticity we performed rheological experiments with Dictyostelium discoideum wild-type cells (AX2) and mutant cells altered by homologous recombination to lack alpha-actinin (AHR), the ABP120 gelation factor (GHR), or both of these F-actin cross-linking proteins (AGHR). Oscillatory and steady flow measurements of Dictyostelium wild-type cells in a torsion pendulum showed that there is a large elastic component to the viscoelasticity of the cell pellet. Quantitative rheological measurements were performed with an electronic plate-and-cone rheometer, which allowed determination of G', the storage shear modulus, and G", the viscous loss modulus, as a function of time, frequency, and strain, respectively. Whole cell viscoelasticity depends strongly on all three parameters, and comparison of wild-type and mutant strains under identical conditions generally produced significant differences. Especially stress relaxation experiments consistently revealed a clear difference between cells that lacked alpha-actinin as compared with wild-type cells or transformants without ABP120 gelation factor, indicating that alpha-actinin plays an important role in cell elasticity. Direct observation of cells undergoing shear deformation was done by incorporating a small number of AX2 cells expressing the green fluorescent protein of Aequorea victoria and visualizing the strained cell pellet by fluorescence and phase contrast microscopy. These observations confirmed that the shear strain imposed by the rheometer does not injure the cells and that the viscoelastic response of the cell pellet is due to deformation of individual cells.  相似文献   

8.
The Ca2+-sensitive actin-binding protein isolated from Dictyostelium discoideum, 30,000-D protein (Fechheimer and Taylor: J. Biol. Chem. 259:4514-4520, 1984;) has recently been localized in filipodia of substrate-adhered amoebae (Fechheimer: J. Cell Biol. 104:1539-1551, 1987). We have determined that this protein has a Mr of 34,000 daltons and is strictly colocalized with actin filaments in both substrate-attached Dictyostelium amoebae and cultured fibroblasts. 3T3 fibroblasts, as well as normal and virally transformed rat kidney fibroblasts (NRK) contain a 34-kilodalton (kD) protein that cross-reacts specifically with antibody to the Dictyostelium bundling protein. Mammalian 34-kD protein is colocalized with F-actin in stress fibers and the cortical cytoskeleton in substrate-adhered fibroblasts. In substrate-adhered vegetative Dictyostelium, F-actin and 34-kD protein are concentrated in regions of the cell cortex exhibiting filipodia and membrane ridges. Multiple filipodia formed after exposure to the chemoattractant folic acid stain intensely for 34-kD protein, implying participation in the assembly of actin bundles during filipod formation. The cortex of pseudopodia also contained high concentrations of bundling protein, but pseudopod interiors did not. In contrast to vegetative Dictyostelium, F-actin and 34-kD protein were not colocalized in cells that had progressed through the developmental cycle. In fruiting bodies, 34-kD protein was detected by immunofluorescence microscopy only in prespore cells, while F-actin appeared in stalk cells and spores.  相似文献   

9.
A protein purified from cytoskeletal fractions of Dictyostelium discoideum proved to be a member of the fimbrin/plastin family of actin-bundling proteins. Like other family members, this Ca(2+)-inhibited 67-kDa protein contains two EF hands followed by two actin-binding sites of the alpha-actinin/beta-spectrin type. Dd plastin interacted selectively with actin isoforms: it bound to D. discoideum actin and to beta/gamma-actin from bovine spleen but not to alpha-actin from rabbit skeletal muscle. Immunofluorescence labeling of growth phase cells showed accumulation of Dd plastin in cortical structures associated with cell surface extensions. In the elongated, streaming cells of the early aggregation stage, Dd plastin was enriched in the front regions. To examine how the bundled actin filaments behave in myosin II-driven motility, complexes of F-actin and Dd plastin were bound to immobilized heavy meromyosin, and motility was started by photoactivating caged ATP. Actin filaments were immediately propelled out of bundles or even larger aggregates and moved on the myosin as separate filaments. This result shows that myosin can disperse an actin network when it acts as a motor and sheds light on the dynamics of protein-protein interactions in the cortex of a motile cell where myosin II and Dd plastin are simultaneously present.  相似文献   

10.
We have studied the abundance, relative gelation activity, and distribution of the 95,000-dalton actin-binding protein in Dictyostelium discoideum amoebae. The 95,000-dalton protein was a prominent polypeptide as assessed using quantitative densitometry and radioimmunoassay. We estimated that this protein comprised approximately 1.2% of the protein in a soluble extract of amoebae. The molar ratio of the dimeric 95,000-dalton protein to actin in the soluble extract was 1:30. The apparent viscosities of actin mixtures with either the purified 95,000-dalton protein or the soluble extract were measured by falling ball viscometry in an attempt to assess the contribution of the 95,000-dalton protein to gelation of the soluble extract. The gelation of the soluble extract was significantly less than that expected from the contribution of the 95,000-dalton protein alone. Consequently, we questioned the validity of quantitative analyses of the contributions of specific actin-binding proteins to the gelation of cell extracts. The apparent distribution of the 95,000- dalton protein was observed in chemically fixed and extracted cells by immunofluorescence microscopy and compared with the distribution of cytoplasm and organelles visible using light microscopy. The 95,000- dalton protein was dispersed throughout the cytoplasm of fixed cells, was apparently excluded from prominent organelles, and displayed brightest fluorescence in regions of hyaline cytoplasm. These regions of hyaline cytoplasm that exhibited the brightest fluorescence were observed in the cortical region of rounded cells and in pseudopods of polarized cells. Thus, cell shape and polarity may also have influenced the apparent distribution of the 95,000-dalton protein observed by immunofluorescence microscopy. Study of the distribution of fluorescein- labeled ovalbumin injected into living cells supported the interpretation that the thickness of the cell and the distribution of organelles contributed to the apparent distribution of the 95,000- dalton protein observed in fixed cells using immunofluorescence microscopy. We suggest that the 95,000-dalton protein contributes to modulation of the consistency and contractility of the cytoplasm of D. discoideum amoebae, since it could cross-link actin filaments in vitro in a reversible process that was regulated by changes in the concentration of calcium and of protons, and since it was present in large quantity in the cytoplasm of these cells.  相似文献   

11.
A calcium-sensitive actin-binding protein having a subunit molecular mass of 30,000 daltons (30K protein) has been isolated from Dictyostelium discoideum. Structural, immunological, and functional analyses demonstrated that the 30K protein was distinct from other actin-binding proteins of D. discoideum. A native molecular mass of 31,700 daltons was determined by equilibrium sedimentation, indicating that the protein is monomeric. The Stokes radius was 30 A. The frictional coefficient calculated from these measurements was 1.44, indicating an asymmetric shape. The 30K protein induced an increase in the viscosity of a solution of F-actin. Bundles of actin filaments were observed in negatively stained mixtures of actin and the 30K protein. Both the formation of filament bundles and the increases in viscosity of actin induced by the 30K protein were observed in the presence of 1 X 10(-8) M but not 2 X 10(-6) M calcium. Variation of the pH from 6.6 to 7.8 had no effect on the activity of the 30K protein. Calcium induced neither a large change in quaternary structure of the 30K protein nor a restriction of the lengths of actin filaments by the 30K protein. The apparent affinity of the 30K protein for actin was decreased in the presence of calcium. Reversible cross-linking of actin filaments by the 30K protein may contribute to regulation of the consistency and contractility of cytoplasm in D. discoideum.  相似文献   

12.
Vibrio cholerae is a Gram-negative bacterial pathogen that exports enterotoxins, which alter host cells through a number of mechanisms resulting in diarrheal disease. Among the secreted toxins is the multifunctional, autoprocessing RTX toxin (MARTX(Vc)), which disrupts actin cytoskeleton by covalently cross-linking actin monomers into oligomers. The region of the toxin responsible for cross-linking activity is the actin cross-linking domain (ACD). In this study, we demonstrate unambiguously that ACD utilizes G- and not F-actin as a substrate for the cross-linking reaction and hydrolyzes one molecule of ATP per cross-linking event. Furthermore, major actin-binding proteins that regulate actin cytoskeleton in vivo do not block the cross-linking reaction in vitro. Cofilin inhibits the cross-linking of G- and F-actin, at a high mole ratio to actin but accelerates F-actin cross-linking at low mole ratios. DNase I completely blocks the cross-linking of actin, likely due to steric hindrance with one of the cross-linking sites on actin. In the context of the holotoxin, the inhibition of Rho by the Rho-inactivating domain of MARTX(Vc) (Sheahan, K. L., and Satchell, K. J. F. (2007) Cell. Microbiol. 9, 1324-1335) would accelerate F-actin depolymerization and provide G-actin, alone or in complex with actin-binding proteins, for cross-linking by ACD, ultimately leading to the observed rapid cell rounding.  相似文献   

13.
We report studies of the fission yeast fimbrin-like protein Fim1, which contains two EF-hand domains and two actin-binding domains (ABD1 and ABD2). Fim1 is a component of both F-actin patches and the F-actin ring, but not of F-actin cables. Fim1 cross-links F-actin in vitro, but a Fim1 protein lacking either EF-hand domains (Fim1A12) or both the EF-hand domains and ABD1 (Fim1A2) has no actin cross-linking activity. Overexpression of Fim1 induced the formation of F-actin patches throughout the cell cortex, whereas the F-actin patches disappear in cells overexpressing Fim1A12 or Fim1A2. Thus, the actin cross-linking activity of Fim1 is probably important for the formation of F-actin patches. The overexpression of Fim1 also excluded the actin-depolymerizing factor Adf1 from the F-actin patches and inhibited the turnover of actin in these structures. Thus, Fim1 may function in stabilizing the F-actin patches. We also isolated the gene encoding Acp1, a subunit of the heterodimeric F-actin capping protein. fim1 acp1 double null cells showed more severe defects in the organization of the actin cytoskeleton than those seen in each single mutant. Thus, Fim1 and Acp1 may function in a similar manner in the organization of the actin cytoskeleton. Finally, genetic studies suggested that Fim1 may function in cytokinesis in cooperation with Cdc15 (PSTPIP) and Rng2 (IQGAP), respectively.  相似文献   

14.
Thyroid spectrin has been obtained pure from pig thyroid glands. This protein, composed of two non-identical polypeptide chains of 240 kDa and 235 kDa, appears to possess the same structural and immunological properties as well as the same calmodulin and actin-binding properties as brain spectrin. Through cross-linking of actin filaments it is a potent gelation factor for F-actin solutions. It represents one of the major protein of the cytoskeleton underlying the thyroid plasma membrane together with myosin, alpha-actinin and actin.  相似文献   

15.
Chromaffin cells, secretory cells of the adrenal medulla, have been shown to contain actin and other contractile proteins, which might be involved in the secretory process. Actin and Ca++-sensitive actin-binding proteins were purified from bovine adrenal medulla on affinity columns using DNase-I as a ligand. Buffers that contained decreasing Ca++ concentrations were used to elute three major proteins of 93, 91, and 85 kD. The bulk of the actin was eluted with guanidine-HCl buffer plus some 93- and 91-kD proteins. These Ca++-sensitive regulatory proteins were shown to inhibit the gelation of actin using the low-shear falling ball viscometer and by electron microscopy. Actin filaments were found to be shortened by fragmentation. Using antibody raised against rabbit lung macrophage gelsolin, proteolytic digestion with Staphylococcus V8 protease and two-dimensional gel electrophoresis, the 91-kD actin-binding protein was shown to be a gelsolin-like protein. The 93-kD actin-binding protein also showed cross-reactivity with anti-gelsolin antibody, similar peptide maps, and a basic-shift in pHi indicating that this 93-kD protein is a brevin-like protein, derived from blood present abundantly in adrenal medulla. Purification from isolated chromaffin cells demonstrated the presence of 91- and 85-kD proteins, whereas the 93-kD protein was hardly detectable. The 85-kD protein is not a breakdown product of brevin-like or gelsolin-like proteins. It did not cross-react with anti-gelsolin antibody and showed a very different peptide map after mild digestion with V8 protease. Antibodies were raised against the 93- and 91-kD actin-binding proteins and the 85-kD actin-binding protein. Antibody against the 85-kD protein did not cross-react with 93- and 91-kD proteins and vice versa. In vivo, the cytoskeleton organization of chromaffin secretory cells is not known, but appears to be under the control of the intracellular concentration of free calcium. The ability of calcium to activate the gelsolin-like protein, and as shown elsewhere to alter fodrin localization, provides a mechanism for gel-sol transition that might be essential for granule movement and membrane-membrane interactions involved in the secretory process.  相似文献   

16.
Gelation factor (ABP120) is one of the principal actin-cross-linking proteins of Dictyostelium discoideum. The extended molecule has an N-terminal 250-residue actin-binding domain and a rod constructed from six 100-residue repeats that have an Ig fold. The ability to dimerize is crucial to the actin cross-linking function of gelation factor and is mediated by the rod in which the two chains are arranged in an antiparallel fashion. We report the 2.2 A resolution crystal structure of rod domains 5 and 6, which shows that dimerization is mediated primarily by rod domain 6 and is the result of a double edge-to-edge extension of beta-sheets. Thus, contrary to earlier proposals, the chains of the dimeric gelation factor molecule overlap only within domain 6, and domains 1-5 do not pair with domains from the other chain. This information allows construction of a model of the gelation factor molecule and suggests how the chains in the related molecule filamin (ABP280) may interact.  相似文献   

17.
Cofilin/ADF proteins are a ubiquitously expressed family of F-actin depolymerizing factors found in eukaryotic cells including plants. In vitro, cofilin/ADF activity has been shown to be essential for actin driven motility, by accelerating actin filament turnover. Three actin depolymerizing factors (n-cofilin, m-cofilin, ADF) can be found in mouse and human. Here we show that in mouse the non-muscle-specific gene-n-cofilin-is essential for migration of neural crest cells as well as other cell types in the paraxial mesoderm. The main defects observed in n-cofilin mutant embryos are an impaired delamination and migration of neural crest cells, affecting the development of neural crest derived tissues. Neural crest cells lacking n-cofilin do not polarize, and F-actin bundles or fibers are not detectable. In addition, n-cofilin is required for neuronal precursor cell proliferation and scattering. These defects result in a complete lack of neural tube closure in n-cofilin mutant embryos. Although ADF is overexpressed in mutant embryos, this cannot compensate the lack of n-cofilin, suggesting that they might have a different function in embryonic development. Our data suggest that in mammalian development, regulation of the actin cytoskeleton by the F-actin depolymerizing factor n-cofilin is critical for epithelial-mesenchymal type of cell shape changes as well as cell proliferation.  相似文献   

18.
To define the actin-binding site within the NH2-terminal domain (residues 1-245) of chick smooth muscle alpha-actinin, we expressed a series of alpha-actinin deletion mutants in monkey Cos cells. Mutant alpha-actinins in which residues 2-19, 217-242, and 196-242 were deleted still retained the ability to target to actin filaments and filament ends, suggesting that the actin-binding site is located within residues 20-195. When a truncated alpha-actinin (residues 1-290) was expressed in Cos cells, the protein localized exclusively to filament ends. This activity was retained by a deletion mutant lacking residues 196-242, confirming that these are not essential for actin binding. The actin-binding site in alpha-actinin was further defined by expressing both wild-type and mutant actin-binding domains as fusion proteins in E. coli. Analysis of the ability of such proteins to bind to F-actin in vitro showed that the binding site was located between residues 108 and 189. Using both in vivo and in vitro assays, we have also shown that the sequence KTFT, which is conserved in several members of the alpha-actinin family of actin-binding proteins (residues 36-39 in the chick smooth muscle protein) is not essential for actin binding. Finally, we have established that the NH2-terminal domain of dystrophin is functionally as well as structurally homologous to that in alpha-actinin. Thus, a chimeric protein containing the NH2-terminal region of dystrophin (residues 1-233) fused to alpha-actinin residues 244-888 localized to actin-containing structures when expressed in Cos cells. Furthermore, an E. coli-expressed fusion protein containing dystrophin residues 1-233 was able to bind to F-actin in vitro.  相似文献   

19.
Triton-insoluble cytoskeletons were isolated from Dictyostelium discoideum AX3 cells prior to and following stimulation with 2'deoxy cyclic adenosine monophosphate (cAMP). Temporal changes in the content of actin and a 120,000 dalton actin-binding protein (ABP-120) in cytoskeletons following stimulation were monitored. Both actin and ABP-120 were incorporated into the cytoskeleton at 30-40 seconds following stimulation, which is cotemporal with the onset of pseudopod extension during stimulation of amoebae with chemoattractants. Changes in the content of total cytoskeletal protein and cytoskeletal myosin were determined under the same experimental conditions as controls. These proteins exhibited different kinetics from those of cytoskeletal ABP-120 and actin following the addition of 2'deoxy cAMP. The authors concluded that the association of ABP-120 with the cytoskeleton is regulated during cAMP signalling. Furthermore, these results indicate that ABP-120 is involved in cross-linking newly assembled actin filaments into the cytoskeleton during chemoattractant-stimulated pseudopod extension.  相似文献   

20.
While actin polymerization and depolymerization are both essential for cell movement, few studies have focused on actin depolymerization. In vivo, depolymerization can occur exceedingly rapidly and in a spatially defined manner: the F-actin in the lamellipodia depolymerizes in 30 s after chemoattractant removal (Cassimeris, L., H. McNeill, and S. H. Zigmond. 1990. J. Cell Biol. 110:1067-1075). To begin to understand the regulation of F-actin depolymerization, we have examined F-actin depolymerization in lysates of polymorphonuclear leukocytes (PMNs). Surprisingly, much of the cell F-actin, measured with a TRITC-phalloidin-binding assay, was stable after lysis in a physiological salt buffer (0.15 M KCl): approximately 50% of the F-actin did not depolymerize even after 18 h. This stable F-actin included lamellar F-actin which could still be visualized one hour after lysis by staining with TRITC-phalloidin and by EM. We investigated the basis for this stability. In lysates with cell concentrations greater than 10(7) cells/ml, sufficient globular actin (G-actin) was present to result in a net increase in F-actin. However, the F-actin stability was not solely because of the presence of free G-actin since addition of DNase I to the lysate did not increase the F-actin loss. Nor did it appear to be because of barbed end capping factors since cell lysates provided sites for barbed end polymerization of exogenous added actin. The stable F-actin existed in a macromolecular complex that pelleted at low gravitational forces. Increasing the salt concentration of the lysis buffer decreased the amount of F-actin that pelleted at low gravitational forces and increased the amount of F-actin that depolymerized. Various actin-binding and cross-linking proteins such as tropomyosin, alpha-actinin, and actin-binding protein pelleted with the stable F-actin. In addition, we found that alpha-actinin, a filament cross-linking protein, inhibited the rate of pyrenyl F-actin depolymerization. These results suggested that actin cross-linking proteins may contribute to the stability of cellular actin after lysis. The activity of crosslinkers may be regulated in vivo to allow rapid turnover of lamellipodia F-actin.  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