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.
Actin exists as a monomer (G-actin) which can be polymerized to filaments) F-actin) that under the influence of actin-binding proteins and polycations bundle and contribute to the formation of the cytoskeleton. Bundled actin from lysed cells increases the viscosity of sputum in lungs of cystic fibrosis patients. The human host defense peptide LL-37 was previously shown to induce actin bundling and was thus hypothesized to contribute to the pathogenicity of this disease. In this work, interactions between actin and the cationic LL-37 were studied by optical, proteolytic and surface plasmon resonance methods and compared to those obtained with scrambled LL-37 and with the cationic protein lysozyme. We show that LL-37 binds strongly to CaATP-G-actin while scrambled LL-37 does not. While LL-37, at superstoichiometric LL-37/actin concentrations polymerizes MgATP-G-actin, at lower non-polymerizing concentrations LL-37 inhibits actin polymerization by MgCl2 or NaCl. LL-37 bundles Mg-F-actin filaments both at low and physiological ionic strength when in equimolar or higher concentrations than those of actin. The LL-37 induced bundles are significantly less sensitive to increase in ionic strength than those induced by scrambled LL-37 and lysozyme. LL-37 in concentrations lower than those needed for actin polymerization or bundling, accelerates cleavage of both monomer and polymer actin by subtilisin. Our results indicate that the LL-37-actin interaction is partially electrostatic and partially hydrophobic and that a specific actin binding sequence in the peptide is responsible for the hydrophobic interaction. LL-37-induced bundles, which may contribute to the accumulation of sputum in cystic fibrosis, are dissociated very efficiently by DNase-1 and also by cofilin.  相似文献   

2.
Three-dimensional structures of actin bundles formed with polycations were observed by using transmission electron microtomography and atomic force microscopy. We found, for the first time, that the cross-sectional morphology of actin bundles depends on the polycation species and ionic strength, while it is insensitive to the degree of polymerization and concentration of polycation. Actin bundles formed with poly-N-[3-(dimethylamino)propyl] acrylamide methyl chloride quaternary show a ribbon-like cross-sectional morphology in low salt concentrations that changes to cylindrical cross-sectional morphology with hexagonal packing of the actin filaments in high salt concentrations. Contrastingly, actin bundles formed with poly-L-lysine show triangular cross-sectional morphology with hexagonal packing of the actin filaments. These variations in cross-sectional morphology are discussed in terms of anisotropy in the electrostatic energy barrier.  相似文献   

3.
Summary Parallel bundles of actin filaments at the cortex-endoplasm interface provide tracks for myosin-generated cytoplasmic streaming in characean internodes. These bundles resist disassembly or structural modification when exposed to 10 μM cytochalasin D (CD) even though this concentration of CD rapidly (within minutes) but reversibly arrests streaming. Unexpectedly, we discovered that prolonged treatment with lower concentrations of CD could partially disassemble the subcortical actin bundles. Actin bundles became discontinuous following one- to several-day treatment with concentrations (6 μM) that reduced but did not arrest streaming, and the residual fragments mostly remained parallel to the chloroplast files. When microtubules were concurrently disassembled with tubulin-specific drugs, however, low CD concentrations (2.5–3 μM) completely arrested bulk streaming, disrupted the largely 2-dimensional actin bundle array and caused the formation of a coarse, thick-meshed actin network that extended from the cortex to the endoplasm. Despite such massive reconstruction, drug removal enabled cells to recover continuous parallel bundles and streaming. Recovery was possible if both or just one of the drugs were removed. In recovered cells, the streaming pattern frequently redeveloped in new directions that did not follow the chloroplast files, and later, chloroplast files readjusted to the new polarity established by the actin bundles. This first report on the complete and reversible disassembly of characean actin bundles provides new insights into the mechanism of actin bundle assembly and organization and supports the idea of indirect interactions between actin filaments and microtubules.  相似文献   

4.
Talin is an adhesion plaque protein believed important in linking actin filaments to the plasma membrane. The nature of a direct talin-actin interaction, however, is complex and has remained unclear. We have systematically characterized the effects of pH, ionic strength, temperature, and protein molar ratio on the interaction between highly purified talin and actin. The ability of talin to increase viscosity of F-actin at 25 degrees C and low ionic strength increased with decreasing pH from 7.3 to 6.4 and increasing molar ratio of talin to actin. At pH 6.4 and low ionic strength, talin could extensively crosslink actin filaments into ordered bundles as shown by negative staining and could cosediment with F-actin at molar ratios as high as one talin to two actin monomers. Talin crosslinked prepolymerized actin filaments to a similar extent as actin filaments polymerized in its presence. The 190-kDa calpain-generated proteolytic fragment of talin bound poorly to actin under conditions favorable for intact talin, but was able to crosslink actin filaments at a lower pH. Increasing the ionic strength within a relatively narrow range significantly decreased ability of talin to bind to actin, regardless of pH. The effects of pH and ionic strength on the talin-actin interaction were rapid and reversible. Low-shear-viscosity studies revealed a strong temperature dependence in the talin-actin interaction with significant crosslinking activity at physiological-like ionic conditions and temperature (37 degrees C). Our results consistently demonstrated that talin crosslinks actin filaments and that this direct interaction is highly sensitive to, and dependent upon, ionic conditions and temperature.  相似文献   

5.
Self-assembly of complex structures is commonplace in biology but often poorly understood. In the case of the actin cytoskeleton, a great deal is known about the components that include higher order structures, such as lamellar meshes, filopodial bundles, and stress fibers. Each of these cytoskeletal structures contains actin filaments and cross-linking proteins, but the role of cross-linking proteins in the initial steps of structure formation has not been clearly elucidated. We employ an optical trapping assay to investigate the behaviors of two actin cross-linking proteins, fascin and α-actinin, during the first steps of structure assembly. Here, we show that these proteins have distinct binding characteristics that cause them to recognize and cross-link filaments that are arranged with specific geometries. α-Actinin is a promiscuous cross-linker, linking filaments over all angles. It retains this flexibility after cross-links are formed, maintaining a connection even when the link is rotated. Conversely, fascin is extremely selective, only cross-linking filaments in a parallel orientation. Surprisingly, bundles formed by either protein are extremely stable, persisting for over 0.5 h in a continuous wash. However, using fluorescence recovery after photobleaching and fluorescence decay experiments, we find that the stable fascin population can be rapidly competed away by free fascin. We present a simple avidity model for this cross-link dissociation behavior. Together, these results place constraints on how cytoskeletal structures assemble, organize, and disassemble in vivo.  相似文献   

6.
Interaction of smooth muscle calponin with 90 kDa heat shock protein (hsp90) was analyzed by means of native gel electrophoresis and affinity chromatography. Under conditions used, calponin and hsp90 form a complex with an apparent dissociation constant in the micromolar range. The major hsp90-binding site is located in the N-terminal (residues 7-144) part of calponin. Addition of calponin to actin-tropomyosin complex results in formation of actin bundles. Hsp90 partially prevents bundle formation without affecting the molar ratio calponin/actin in single actin filaments or actin bundles. At low ionic strength, calponin induces polymerization of G-actin. Hsp90 decreases calponin-induced polymerization of G-actin. It is supposed that hsp90 may be involved in the assembly of actin filaments.  相似文献   

7.
The actin depolymerizing factor (ADF)/cofilin family of proteins interact with actin monomers and filaments in a pH-sensitive manner. When ADF/cofilin binds F-actin it induces a change in the helical twist and fragmentation; it also accelerates the dissociation of subunits from the pointed ends of filaments, thereby increasing treadmilling or depolymerization. Using site-directed mutagenesis we characterized the two actin-binding sites on human cofilin. One target site was chosen because we previously showed that the villin head piece competes with ADF for binding to F-actin. Limited sequence homology between ADF/cofilin and the part of the villin headpiece essential for actin binding suggested an actin-binding site on cofilin involving a structural loop at the opposite end of the molecule to the alpha-helix already implicated in actin binding. Binding through the alpha-helix is primarily to monomeric actin, whereas the loop region is specifically involved in filament association. We have characterized the actin binding properties of each site independently of the other. Mutation of a single lysine residue in the loop region abolishes binding to filaments, but not to monomers. Using the mutation analogous to the phosphorylated form of cofilin (S3D), we show that filament binding is inhibited at physiological ionic strength but not under low salt conditions. At low ionic strength, this mutant induces both the twist change and fragmentation characteristic of wild-type cofilin, but does not activate subunit dissociation. The results suggest a two-site binding to filaments, initiated by association through the loop site, followed by interaction with the adjacent subunit through the "helix" site at the opposite end of the molecule. Together, these interactions induce twist and fragmentation of filaments, but the twist change itself is not responsible for the enhanced rate of actin subunit release from filaments.  相似文献   

8.
Fast actin depolymerization is necessary for cells to rapidly reorganize actin filament networks. Utilizing a Listeria fluorescent actin comet tail assay to monitor actin disassembly rates, we observed that although a mixture of actin disassembly factors (cofilin, coronin, and actin-interacting protein 1 is sufficient to disassemble actin comet tails in the presence of physiological G-actin concentrations this mixture was insufficient to disassemble actin comet tails in the presence of physiological F-actin concentrations. Using biochemical complementation, we purified cyclase-associated protein (CAP) from thymus extracts as a factor that protects against the inhibition of excess F-actin. CAP has been shown to participate in actin dynamics but has been thought to act by liberating cofilin from ADP·G-actin monomers to restore cofilin activity. However, we found that CAP augments cofilin-mediated disassembly by accelerating the rate of cofilin-mediated severing. We also demonstrated that CAP acts directly on F-actin and severs actin filaments at acidic, but not neutral, pH. At the neutral pH characteristic of cytosol in most mammalian cells, we demonstrated that neither CAP nor cofilin are capable of severing actin filaments. However, the combination of CAP and cofilin rapidly severed actin at all pH values across the physiological range. Therefore, our results reveal a new function for CAP in accelerating cofilin-mediated actin filament severing and provide a mechanism through which cells can maintain high actin turnover rates without having to alkalinize cytosol, which would affect many biochemical reactions beyond actin depolymerization.  相似文献   

9.
We have examined the structure of actin-binding molecules in solution and interacting with actin filaments. At physiological ionic strength, actin-binding protein has a Mr value of 540 × 103 as determined by direct and indirect hydrodynamic measurements. It is an asymmetrical dimer composed of 270 × 103 dalton subunits. Viewed in the electron microscope after negative staining or low angle shadowing, actin-binding protein molecules assume a broad range of conformations varying from closed circular structures to fully extended strands 162 nm in contour length. All configurations are apparently derived from the same structure which consists of two monomer chains connected end-to-end. The radius of gyration determined from the electron microscopic images was 21.3 nm in agreement with the value of 17.6 nm calculated from hydrodynamic assays. The average axial ratio from hydrodynamic measurements was 17:1, whereas fully extended dimer molecules in the electron microscope would have an axial ratio of 54:1. All of these observations indicate that actin-binding protein dimers are extremely flexible. The flexibility parameter λ (Landau &; Lifshits, 1958) for actinbinding protein is 0.18 nm?1.As determined by sedimentation, actin-binding protein binds to actin filaments with a Ka value of 2 × 106m?1 and a capacity of one dimer to 14 actin monomers in filaments. After incubation of high concentrations (molar ratio to actin ≥ 1:10) of actin-binding protein with actin filaments, long filament bundles are visible in the electron microscope. Under these conditions, actin-binding protein molecules decorate the actin filaments in the bundles at regular 40 nm intervals or once every 15 monomers, approximately equivalent to the binding capacity measured by sedimentation. Low concentrations of actin-binding protein (molar ratio to actin ≥ 1:50) which promote the gelation of actin filaments in solution, did not detectably alter the isotropy of the actin filaments. Direct visualization of actinbinding protein molecules between actin filaments in the electron microscope showed that dimers are sufficient for crossbridging of actin filaments and that actinbinding protein dimers are bipolar, composed of monomers connected head-to-head and having actin-binding sites located on the free tails.We conclude that actin-binding protein is a dimer at physiological ionic strength. Each dimer has two actin filament binding sites and is therefore sufficient to gel actin filaments in solution. The length and flexibility of the actin-binding protein subunits render this molecule structurally suited for the crosslinking of large helical filaments into isotropic networks.  相似文献   

10.
The detailed substructure of actin filament bundles in microvilli of fertilized sea urchin eggs has been studied by analysing electron microscope images of negatively stained specimens. Transverse stripes which repeat about every 130 Å along the axis of a bundle, as previously observed by Burgess & Schroeder (1977), reflect the positions of cross-bridges that connect the filaments into a bundle. Analysis of optical transforms of the micrographs reveals that there are approximately 14 actin monomers between cross-overs of the two long-pitch helical strands of the actin filaments, with three cross-bridges in this interval. The structure is basically similar to that of the hexagonally packed bundles prepared in vitro from high speed supernatants of sea urchin eggs by Kane (1975) and analyzed by DeRosier et al. (1977). One clear difference, however, is that the in vivo microvillar filament bundles are supercoiled, giving rise to long axial repeats of 1500 to 2000 Å.Computationally filtered images of regions that were only slightly supercoiled reveal the relative alignment of filaments within the bundles and show that crossbridges appear to interact with four actin monomers, apparently linking two actin monomers on one strand of one filament to the nearest two monomers on a neighbouring filament. However, the cross-bridges are not spaced at equal intervals corresponding to four actin subunits, presumably because of the lack of hexagonal symmetry in the individual filaments, which have about 14 actin monomers between cross-overs. Instead, the cross-bridges are arranged quasiequivalently along the longitudinal axis of the bundles, in steps of four or five actin subunit spacings (28 Å each).  相似文献   

11.
Both the sliding velocity of fluorescently labeled actin filament and its persistence length as an index of the bending flexibility of the filament were examined in the motility assay as varying the pH values of the solution for preparing actin filaments. When the pH value was varied from 5.0 to 9.0 in the solution in which actin filaments were formed from the constituent monomers, the motile performance of Mg2+ bound actin filaments (Mg-F-actin) was apparently suppressed compared to the case of Ca2+ bound ones (Ca-F-actin). The persistence length for Ca-F-actin gradually increased with the increase of the pH value while the similar length for Mg-F-actin remained rather independent of the value. The largest sliding velocity of the filament, on the other hand, obtained at the persistence length of roughly 6 μm for both cases of Mg-F-actin and Ca-F-actin.  相似文献   

12.
Small-angle X-ray scattering was used to probe the structure of actin in the presence of cryosolvents: 1,2-propanediol, glycerol, or a mixture of both solvents. In media devoid of polymerizing salts, a radius of gyration of 23 Å is measured, as expected from the literature. In the presence of 1,2-propanediol alone, the scattering pattern begins to exhibit the characteristic slope of elongated objects with a non-negligible thickness, such as actin filaments polymerized in 40 mM KCl and 1 mM MgCl2. However, only short fragments (radius of gyration 40 Å) are generated. We infer that in a medium of low ionic strength containing 15% 1,2-propanediol, actin assumes a structure closer to that of filamentous actin. 1,2-propanediol apparently induces nucleation of oligomers, as with polymerizing salts, but no propagation occurs. Glycerol and/or propanediol induce no alteration in the structure of individual salt-polymerized actin filaments. Aggregation occurs with propanediol, even in the presence of glycerol. Glycerol alone has no such effect. No shortening is detected within the scale covered, with either solvent, although 1,2-propanediol is known to shorten actin filaments. We suggest that in the absence of salts, 1,2-propanediol induces a conformational change in monomeric actin that is necessary for nucleation. This could correlate with a conformational change of actin protomers within microfilaments observed in the presence of 1,2-propanediol by other authors using different techniques.Abbreviations SAXS small-angle X-ray scattering - G-actin globular monomeric actin - F-actin filamentous polymerized actin Correspondence to: E. Pajot-Augy  相似文献   

13.
We have used two in vitro motility assays to study the relative movement of actin and myosin from turkey gizzards (smooth muscle) and human platelets. In the Nitella-based in vitro motility assay, myosin-coated polymer beads move over a fixed substratum of actin bundles derived from dissection of the alga, Nitella, whereas in the sliding actin filament assay fluorescently labeled actin filaments slide over myosin molecules adhered to a glass surface. Both assay systems yielded similar relative velocities using smooth muscle myosin and actin under our standard conditions. We have studied the effects of ATP, ionic strength, magnesium, and tropomyosin on the velocity and found that with the exception of the dependence on MgCl2, the two assays gave very similar results. Calcium over a concentration of pCa 8 to 4 had no effect on the velocity of actin filaments. Phosphorylated smooth muscle myosin propelled filaments of smooth muscle and skeletal muscle actin at the same rate. Phosphorylated smooth muscle and cytoplasmic myosin monomers also moved actin filaments, demonstrating that filament formation is not required for movement.  相似文献   

14.
The thermodynamic properties of the actin filaments prepared from cardiomyocytes were investigated with differential scanning calorimetry. This method could distinguish between the α-cardiac and α-skeletal components of the actin filaments polymerised from ADP-actin monomers by their different melting temperatures (Tm). Similar separation was not possible with filaments polymerised from ATP-actin monomers. Further analyses revealed that the activation energy (Eact) was greater for filaments of α-skeletal actin than for α-cardiac actin monomers when the filaments were polymerised from ADP-actin monomers. These results showed that the α-cardiac actin filaments were thermodynamically less stable than the filaments of α-skeletal actin and their difference was nucleotide dependent. Based on these results and considering previous observations it was concluded that the existence of two actin isoforms and their nucleotide dependent conformational differences are part of the tuning regulatory mechanism by which the cardiac muscle cells can maintain their biological function under pathological conditions.  相似文献   

15.
Aip1 (actin interacting protein 1) is ubiquitous in eukaryotic organisms, where it cooperates with cofilin to disassemble actin filaments, but neither its mechanism of action nor its biological functions have been clear. We purified both fission yeast and human Aip1 and investigated their biochemical activities with or without cofilin. Both types of Aip1 bind actin filaments with micromolar affinities and weakly nucleate actin polymerization. Aip1 increases up to 12-fold the rate that high concentrations of yeast or human cofilin sever actin filaments, most likely by competing with cofilin for binding to the side of actin filaments, reducing the occupancy of the filaments by cofilin to a range favorable for severing. Aip1 does not cap the barbed ends of filaments severed by cofilin. Fission yeast lacking Aip1 are viable and assemble cytokinetic contractile rings normally, but rings in these Δaip1 cells accumulate 30% less myosin II. Further, these mutant cells initiate the ingression of cleavage furrows earlier than normal, shortening the stage of cytokinetic ring maturation by 50%. The Δaip1 mutation has negative genetic interactions with deletion mutations of both capping protein subunits and cofilin mutations with severing defects, but no genetic interaction with deletion of coronin.  相似文献   

16.
The high actin-based motility rates observed in nonmuscle cells require the per-second addition of 400-500 monomers to the barbed ends of growing actin filaments. The chief polymerization-competent species is profilin.actin.ATP (present at 5-40 microM intracellular concentrations), whereas G-actin.ATP is much less abundant ( approximately 0.1-1 microM). While earlier studies unambiguously demonstrated that profilin.actin is highly concentrated within the polymerization zone, profilin-actin localization on the motile surface cannot increase the local solution-phase concentration of polymerizable actin. To explain these high rates of actin polymerization, we present and analyze a novel polymerization model in which monomers are directly transferred to growing filament ends in the actoclampin motor. This direct-transfer polymerization mechanism endows the polymerization zone with properties unavailable to bulk-phase actin monomers, and our model also indicates why profilin is the ideal mobile carrier for actin monomers.  相似文献   

17.
The full-length human papillomavirus 16 major capsid protein L1 is expressed in Saccharomyces cerevisiae as virus-like particles (VLPs). However, yeast-expressed human papillomavirus 16 particles are irregular in shape and are prone to aggregate. When disassembled and reassembled, the resulting particles have improved stability and solubility. We have examined VLP dissociation and reassembly to define the important features of the assembly mechanism. We found that the VLPs rapidly disassemble at pH 8.2 and low ionic strength in the presence of low concentrations of reducing agents. The pH dependence of assembly kinetics and extent of assembly under reducing conditions were differentially sensitive to ionic strength. Assembly at pH 5.2 was very fast and led to heavily aggregated particles. This sort of kinetic trap is expected for overinitiated assembly. We observed that reassembly at pH 6.2, 7.2, and 8.2 yielded regular particles over a broad range of ionic strength. At these three pH values, assembly was quantitative at 1 M NaCl. At pH 7.2, much more than at pH 6.2 or pH 8.2, assembly decreased monotonically with ionic strength. The free energy of association ranged from − 8 to − 10 kcal/mol per pentamer. The effect of pH on assembly was further investigated by examining dissociation of reassembled particles. Though indistinguishable by negative stain electron microscopy, particles assembled at pH 7.2 disassembled slower than pH 5.2, 6.2, or 8.2 VLPs. We hypothesize that pH 7.2 assembly reactions lead to formation of particles with conformationally different interactions.  相似文献   

18.
Organization and expression of Drosophila tropomyosin genes   总被引:12,自引:0,他引:12  
It has been shown (Jockusch &; Isenberg, 1981) that vinculin (130K protein) binds to actin and induces actin filaments to form bundles even at low ionic strength. Here, we present structural details on the vinculin molecule itself and on its interaction with actin. In negatively stained preparations, vinculin appeared as a globular protein with an average diameter of 85 Å. The ability of vinculin to form actin filament bundles was confirmed using shadowing techniques and gel analysis of sedimented material. Analysis of vinculin-induced paracrystals by optical diffraction and computer processing revealed their structural similarity to Mg-induced paracrystals. The lateral position of vinculin on surface-exposed actin filaments of such paracrystals was demonstrated directly in electron micrographs and indirectly by labelling vinculin with ferritin-coupled anti-vinculin F(ab′) fragments. Polymerization of actin in the presence of vinculin-coated polystyrene beads did not result in an “end-on” binding of filaments to the beads. Rather, actin bundles were laterally associated with the whole surface of the beads, from where they radiated in a star-like pattern. The growth of actin filaments onto myosin subfragment-I decorated, vinculin-incubated. fixed filament fragments was not inhibited, as was shown directly by electron microscopy and monitored viscometrically in a nucleation assay. These results suggest that in vivo at the site of an adhesion plaque vinculin may link actin filaments together into a suitable configuration to interact with the plasma membrane.  相似文献   

19.
Oxidovanadium(IV), a cationic species (VO2+) of vanadium(IV), binds to several proteins, including actin. Upon titration with oxidovanadium(IV), approximately 100% quenching of the intrinsic fluorescence of monomeric actin purified from rabbit skeletal muscle (G-actin) was observed, with a V50 of 131 μM, whereas for the polymerized form of actin (F-actin) 75% of quenching was obtained and a V50 value of 320 μM. Stern-Volmer plots were used to estimate an oxidovanadium(IV)-actin dissociation constant, with Kd of 8.2 μM and 64.1 μM VOSO4, for G-actin and F-actin, respectively. These studies reveal the presence of a high affinity binding site for oxidovanadium(IV) in actin, producing local conformational changes near the tryptophans most accessible to water in the three-dimensional structure of actin. The actin conformational changes, also confirmed by 1H NMR, are accompanied by changes in G-actin hydrophobic surface, but not in F-actin. The 1H NMR spectra of G-actin treated with oxidovanadium(IV) clearly indicates changes in the resonances ascribed to methyl group and aliphatic regions as well as to aromatics and peptide-bond amide region. In parallel, it was verified that oxidovanadium(IV) prevents the G-actin polymerization into F-actin. In the 0-200 μM range, VOSO4 inhibits 40% of the extent of polymerization with an IC50 of 15.1 μM, whereas 500 μM VOSO4 totally suppresses actin polymerization. The data strongly suggest that oxidovanadium(IV) binds to actin at specific binding sites preventing actin polymerization. By affecting actin structure and function, oxidovanadium(IV) might be responsible for many cellular effects described for vanadium.  相似文献   

20.
In striated muscle the pointed ends of polar actin filaments are directed toward the center of the sarcomer. Formed filaments keep a constant length of about 1 μm. As polymerization and depolymerization at free pointed ends are not sufficiently slow to account for the constant length of the filaments, we searched for proteins which occur in sarcomers and can stabilize the pointed ends of actin filaments. We observed that tropornyosintroponin complex reduces the rate of association and dissociation of actin molecules at the pointed ends more than 30-fold. On the average, every 600 s one association or dissociation reaction has been found to occur at the pointed ends near the critical actin monomer concentration.  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