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.
The Arp2/3 complex, first isolated from Acanthamoeba castellani by affinity chromatography on profilin, consists of seven polypeptides; two actin-related proteins, Arp2 and Arp3; and five apparently novel proteins, p40, p35, p19, p18, and p14 (Machesky et al., 1994). The complex is homogeneous by hydrodynamic criteria with a Stokes' radius of 5.3 nm by gel filtration, sedimentation coefficient of 8.7 S, and molecular mass of 197 kD by analytical ultracentrifugation. The stoichiometry of the subunits is 1:1:1:1:1:1:1, indicating the purified complex contains one copy each of seven polypeptides. In electron micrographs, the complex has a bilobed or horseshoe shape with outer dimensions of approximately 13 x 10 nm, and mathematical models of such a shape and size are consistent with the measured hydrodynamic properties. Chemical cross-linking with a battery of cross-linkers of different spacer arm lengths and chemical reactivities identify the following nearest neighbors within the complex: Arp2 and p40; Arp2 and p35; Arp3 and p35; Arp3 and either p18 or p19; and p19 and p14. By fluorescent antibody staining with anti-p40 and -p35, the complex is concentrated in the cortex of the ameba, especially in linear structures, possibly actin filament bundles, that lie perpendicular to the leading edge. Purified Arp2/3 complex binds actin filaments with a Kd of 2.3 microM and a stoichiometry of approximately one complex molecule per actin monomer. In electron micrographs of negatively stained samples, Arp2/3 complex decorates the sides of actin filaments. EDC/NHS cross-links actin to Arp3, p35, and a low molecular weight subunit, p19, p18, or p14. We propose structural and topological models for the Arp2/3 complex and suggest that affinity for actin filaments accounts for the localization of complex subunits to actin-rich regions of Acanthamoeba.  相似文献   

2.
Profilin, a ubiquitous 12 to 15-kDa protein, serves many functions, including sequestering monomeric actin, accelerating nucleotide exchange on actin monomers, decreasing the critical concentration of the barbed end of actin filaments, and promoting actin polymerization when barbed ends are free. Most previous studies have focused on profilin itself rather than its complex with actin. A high-affinity profilin-actin complex (here called profilactin) can be isolated from a poly-(L)-proline (PLP) column by sequential elution with 3 M and 7 M urea. Profilactin inhibited the elongation rate of pyrenyl-G-actin from filament seeds in a concentration- and time-dependent manner. Much greater inhibition of elongation was observed with spectrin-F-actin than gelsolin-F-actin seeds, suggesting that the major effect of profilactin was due to capping the barbed ends of actin filaments. Its dissociation constant for binding to filament ends was 0.3 microM; the on- and off-rate constants were estimated to be 1.7 x 10(3) M-1 s-1 and 4.5 x 10(-4) s-1, respectively. Purified profilin (obtained by repetitive applications to a PLP column and assessed by silver-stained polyacylamide gels) did not slow the elongation rate of pyrenyl-G-actin from filament seeds. Capping protein could not be detected by Western blotting in the profilactin preparation, but low concentrations of gelsolin did contaminate our preparation. However, prolonged incubation with either calcium or EGTA did not affect capping activity, implying that contaminating gelsolin-actin complexes were not primarily responsible for the observed capping activity. Reapplication of the profilactin preparation to PLP-coupled Sepharose removed both profilin and actin and concurrently eliminated its capping activity. Profilactin that was reapplied to uncoupled Sepharose retained its capping activity. Phosphatidylinositol-4,5-bisphosphate (PIP2) was the most potent phosphoinositol in reducing the capping activity of profilactin. Dissociation of the tight profilactin complex may serve as a unique mechanism by which profilin helps regulate actin filament growth.  相似文献   

3.
Three methods, fluorescence anisotropy of rhodamine-labeled profilin, intrinsic fluorescence and nucleotide exchange, give the same affinity, Kd = 0.1 microM, for Acanthamoeba profilins binding amoeba actin monomers with bound Mg-ATP. Replacement of serine 38 with cysteine created a unique site where labeling with rhodamine did not alter the affinity of profilin for actin. The affinity for rabbit skeletal muscle actin is about 4-fold lower. The affinity for both actins is 5-8-fold lower with ADP bound to actin rather than ATP. Pyrenyliodoacetamide labeling of cysteine 374 of muscle actin reduces the affinity for profilin 10-fold. The affinity of profilin for nucleotide-free actin is approximately 3-fold higher than for Mg-ATP-actin and approximately 24-fold higher than for Mg-ADP-actin. As a result, profilin binding reduces the affinity of actin 3-fold for Mg-ATP and 24-fold for Mg-ADP. Mg-ATP dissociates 8 times faster from actin-profilin than from actin and binds actin-profilin 3 times faster than actin. Mg-ADP dissociates 14 times faster from actin-profilin than from actin and binds actin-profilin half as fast as actin. Thus, profilin promotes the exchange of ADP for ATP. These properties allow profilin to bind a high proportion of unpolymerized ATP-actin in the cell, suppressing spontaneous nucleation but allowing free barbed ends to elongate at more than 500 subunits/second.  相似文献   

4.
Actin filament assembly at the cell surface of the pathogenic bacterium Listeria monocytogenes requires the bacterial ActA surface protein and the host cell Arp2/3 complex. Purified Arp2/3 complex accelerated the nucleation of actin polymerization in vitro, but pure ActA had no effect. However, when combined, the Arp2/3 complex and ActA synergistically stimulated the nucleation of actin filaments. This mechanism of activating the host Arp2/3 complex at the L. monocytogenes surface may be similar to the strategy used by cells to control Arp2/3 complex activity and hence the spatial and temporal distribution of actin polymerization.  相似文献   

5.
The actin-based motility of Listeria monocytogenes requires the addition of actin monomers to the barbed or plus ends of actin filaments. Immunofluorescence micrographs have demonstrated that gelsolin, a protein that both caps barbed ends and severs actin filaments, is concentrated directly behind motile bacteria at the junction between the actin filament rocket tail and the bacterium. In contrast, CapG, a protein that strictly caps actin filaments, fails to localize near intracellular Listeria. To explore the effect of increasing concentrations of gelsolin on bacterial motility, NIH 3T3 fibroblasts stably transfected with gelsolin cDNA were infected with Listeria. The C5 cell line containing 2.25 times control levels of gelsolin supported significantly higher velocities of bacterial movement than did control fibroblasts (mean +/- standard error of the mean, 0.09 +/- 0.003 micro(m)/s [n = 176] versus 0.05 +/- 0.003 micro(m)/s [n = 65]). The rate of disassembly of the Listeria-induced actin filament rocket tail was found to be independent of gelsolin content. Therefore, if increases in gelsolin content result in increases in Listeria-induced rocket tail assembly rates, a positive correlation between gelsolin content and tail length would be expected. BODIPY-phalloidin staining of four different stably transfected NIH 3T3 fibroblast cell lines confirmed this expectation (r = 0.92). Rocket tails were significantly longer in cells with a high gelsolin content. Microinjection of gelsolin 1/2 (consisting of the amino-terminal half of native gelsolin) also increased bacterial velocity by more than 2.2 times. Microinjection of CapG had no effect on bacterial movement. Cultured skin fibroblasts derived from gelsolin-null mice were capable of supporting intracellular Listeria motility at velocities comparable to those supported by wild-type skin fibroblasts. These experiments demonstrated that the surface of Listeria contains a polymerization zone that can block the barbed-end-capping activity of both gelsolin and CapG. The ability of Listeria to uncap actin filaments combined with the severing activity of gelsolin can accelerate actin-based motility. However, gelsolin is not absolutely required for the actin-based intracellular movement of Listeria because its function can be replaced by other actin regulatory proteins in gelsolin-null cells, demonstrating the functional redundancy of the actin system.  相似文献   

6.
Alterations in vascular cell shape and motility occur during developmental processes and in response to injury. Similarly, during tumor vascularization and atherogenesis, endothelial and smooth muscle cells undergo motile and proliferative responses to extracellular cues. Recent inroads into our understanding of signal transduction have identified several candidate pathways by which the extracellular matrix- and growth factor-mediated stimulation of vascular cell motility may be mediated. The multiple and divergent extracellular stimuli that stimulate vascular motile responses may converge on the cytoskeleton via a family of ras-related GTPases. Biochemical analyses as well as examination of cytoskeletal dynamics in vivo indicate that actin polymerization at the forward aspects of spreading cytoplasm is capable of driving forward protrusion formation in the absence of a conventional actin motor. Actin polymerization at the plasma membrane of leading lamellae may be mediated both by de novo nucleation of actin filaments and the generation of free filament ends by uncapping the barbed ends of existing actin filaments. This review summarizes the most recent findings in extracellular-cytoskeletal-signal transduction, therein, providing a framework to explain the remarkable remodeling seen in the vasculature during developmental and disease-related processes.  相似文献   

7.
Structural and functional properties of intrastrand, ANP (N-(4-azido-2-nitrophenyl)-putrescine) cross-linked actin filaments, between Gln-41 and Cys-374 on adjacent monomers, were examined for several preparations of such actin. Extensively cross-linked F-actin (with 12% un-cross-linked monomers) lost at 60 degrees C the ability to activate myosin ATPase at a 100-fold slower rate and unfolded in CD melting experiments at a temperature higher by 11 degrees C than the un-cross-linked actin. Electron microscopy and image reconstruction of these filaments did not reveal any gross changes in F-actin structure but showed a change in the orientation of subdomain 2 and a decrease in interstrand connectivity. Rigor and weak (in the presence of ATP) myosin subfragment (S1) binding and acto-S1 ATPase did not show major changes upon 50% and 90% ANP cross-linking of F-actin; the Kd and Km values were little affected by the cross-linking, and the Vmax decreased by 50% for the extensively cross-linked actin. The cross-linking of actin (50%) decreased the mean speed and the number of sliding filaments in the in vitro motility assays by approximately 35% while the relative force, as measured by using external load in these assays, was inhibited by approximately 25%. The mean speed of actin filaments decreased with the increase in their cross-linking and approached 0 for the 90% cross-linked actin. Also examined were actin filaments reassembled from cross-linked and purified ANP cross-linked dimers, trimers, and oligomers. All of these filaments had the same acto-S1 ATPase and rigor S1 binding properties but different behavior in the in vitro motility assays. Filaments made of cross-linked dimers moved at approximately 50% of the speed of the un-cross-linked actin. The movement of filaments made of cross-linked trimers was inhibited more severely, and the oligomer-made filaments did not move at all. These results show the uncoupling between force generation and other events in actomyosin interactions and emphasize the role of actin filament structure and dynamics in the contractile process.  相似文献   

8.
Adducin is a membrane skeleton protein originally described in human erythrocytes that promotes the binding of spectrin to actin and also binds directly to actin and bundles actin filaments. Adducin is associated with regions of cell-cell contact in nonerythroid cells, where it is believed to play a role in regulating the assembly of the spectrin-actin membrane skeleton. In this study we demonstrate a novel function for adducin; it completely blocks elongation and depolymerization at the barbed (fast growing) ends of actin filaments, thus functioning as a barbed end capping protein (Kcap approximately 100 nM). This barbed end capping activity requires the intact adducin molecule and is not provided by the NH2-terminal globular head domains alone nor by the COOH-terminal extended tail domains, which were previously shown to contain the spectrin-actin binding, calmodulin binding, and phosphorylation sites. A novel difference between adducin and other previously described capping proteins is that it is down-regulated by calmodulin in the presence of calcium. The association of stoichiometric amounts of adducin with the short erythrocyte actin filaments in the membrane skeleton indicates that adducin could be the functional barbed end capper in erythrocytes and play a role in restricting actin filament length. Our experiments also suggest novel possibilities for calcium regulation of actin filament assembly by adducin in erythrocytes and at cell-cell contact sites in nonerythroid cells.  相似文献   

9.
Cdc42, activated with GTPgammaS, induces actin polymerization in supernatants of lysed neutrophils. This polymerization, like that induced by agonists, requires elongation at filament barbed ends. To determine if creation of free barbed ends was sufficient to induce actin polymerization, free barbed ends in the form of spectrin-actin seeds or sheared F-actin filaments were added to cell supernatants. Neither induced polymerization. Furthermore, the presence of spectrin-actin seeds did not increase the rate of Cdc42-induced polymerization, suggesting that the presence of Cdc42 did not facilitate polymerization from spectrin-actin seeds such as might have been the case if Cdc42 inhibited capping or released G-actin from a sequestered pool. Electron microscopy revealed that Cdc42-induced filaments elongated rapidly, achieving a mean length greater than 1 micron in 15 s. The mean length of filaments formed from spectrin-actin seeds was <0.4 micron. Had spectrin-actin seeds elongated at comparable rates before they were capped, they would have induced longer filaments. There was little change in mean length of Cdc42-induced filaments between 15 s and 5 min, suggesting that the increase in F-actin over this time was due to an increase in filament number. These data suggest that Cdc42 induction of actin polymerization requires both creation of free barbed ends and facilitated elongation at these ends.  相似文献   

10.
Regulation of the F-actin severing activity of gelsolin by Ca2+ has been investigated under physiologic ionic conditions. Tryptophan fluorescence intensity measurements indicate that gelsolin contains at least two Ca2+ binding sites with affinities of 2.5 x 10(7) M-1 and 1.5 x 10(5) M-1. At F-actin and gelsolin concentrations in the range of those found intracellularly, gelsolin is able to bind F-actin with half-maximum binding at 0.14 microM free Ca2+ concentration. Steady-state measurements of gelsolin-induced actin depolymerization suggest that half-maximum depolymerization occurs at approximately 0.4 microM free Ca2+ concentration. Dynamic light scattering measurements of the translational diffusion coefficient for actin filaments and nucleated polymerization assays for number concentration of actin filaments both indicate that severing of F-actin occurs slowly at micromolar free Ca2+ concentrations. The data suggest that binding of Ca2+ to the gelsolin-F-actin complex is the rate-limiting step for F-actin severing by gelsolin; this Ca2+ binding event is a committed step that results in a Ca2+ ion bound at a high-affinity, EGTA-resistant site. The very high affinity of gelsolin for the barbed end of an actin filament drives the binding reaction equilibrium toward completion under conditions where the reaction rate is slow.  相似文献   

11.
We have determined diffusion coefficients for small (50- to 70-nm diameter) fluorescein-thiocarbamoyl-labeled Ficoll tracers through F-actin as a function of filament length and cross-linking. fx45 was used to regulate filament length and avidin/biotinylated actin or ABP-280 was used to prepare cross-linked actin gels. We found that tracer diffusion was generally independent of filament length in agreement with theoretical predictions for diffusion through solutions of rods. However, in some experiments diffusion was slower through short (< or = 1.0 micron) filaments, although this result was not consistently reproducible. Measured diffusion coefficients through unregulated F-actin and filaments of lengths > 1.0 micron were more rapid than predicted by theory for tracer diffusion through rigid, random networks, which was consistent with some degree of actin bundling. Avidin-induced cross-linking of biotinylated F-actin did not affect diffusion through unregulated F-actin, but in cases where diffusion was slower through short filaments this cross-linking method resulted in enhanced tracer diffusion rates indistinguishable from unregulated F-actin. This finding, in conjunction with increased turbidity of 1.0-micron filaments upon avidin cross-linking, indicated that this cross-linking method induces F-actin bundling. By contrast, ABP-280 cross-linking retarded diffusion through unregulated F-actin and decreased turbidity. Tracer diffusion under these conditions was well approximated by the diffusion theory. Both cross-linking procedures resulted in gel formation as determined by falling ball viscometry. These results demonstrate that network microscopic geometry is dependent on the cross-linking method, although both methods markedly increase F-actin macroscopic viscosity.  相似文献   

12.
Coronin is a highly conserved actin-associated protein that until now has had unknown biochemical activities. Using microtubule affinity chromatography, we coisolated actin and a homologue of coronin, Crn1p, from Saccharomyces cerevisiae cell extracts. Crn1p is an abundant component of the cortical actin cytoskeleton and binds to F-actin with high affinity (Kd 6 x 10(-9) M). Crn1p promotes the rapid barbed-end assembly of actin filaments and cross-links filaments into bundles and more complex networks, but does not stabilize them. Genetic analyses with a crn1Delta deletion mutation also are consistent with Crn1p regulating filament assembly rather than stability. Filament cross-linking depends on the coiled coil domain of Crn1p, suggesting a requirement for Crn1p dimerization. Assembly-promoting activity is independent of cross-linking and could be due to nucleation and/or accelerated polymerization. Crn1p also binds to microtubules in vitro, and microtubule binding is enhanced by the presence of actin filaments. Microtubule binding is mediated by a region of Crn1p that contains sequences (not found in other coronins) homologous to the microtubule binding region of MAP1B. These activities, considered with microtubule defects observed in crn1Delta cells and in cells overexpressing Crn1p, suggest that Crn1p may provide a functional link between the actin and microtubule cytoskeletons in yeast.  相似文献   

13.
Actin cross-linked between cys 374 and cys 10 via a disulfide-containing bridge, c-A, is completely unpolymerizable even in the presence of phalloidin. Upon the addition of dithiothreitol, c-A polymerizes with high yield, indicating that denaturation due to the modification was almost absent. In the present study we show that cross-linked actin is a useful model for studying the properties of monomeric actin under polymerization conditions. Addition of salt, for example, produced fluorescence changes possibly reflecting conformational transitions but did not lead to the development of phalloidin binding capacity. Cross-linking of the two cysteine residues also caused a decrease in the nucleotide exchange rate by a factor of ca. 3, an effect that was fully reversed by the addition of KCl. Cross-linked actin inhibits DNase I to the same extent as G-actin and binds thymosin beta 4 and profilin as shown by cross-linking studies. Capping capacity for the barbed end of the filament was not observed, although it might have been expected from the fact that both ends of the cross-link are anchored to subdomain 1. Using the 61-FITC derivative of c-A we showed that c-A is able to bind to myosin S1 with a KD in the microM range. In agreement with this, c-A shows actomyosin ATPase activity with a Kapp comparable to that of F-actin, but a Vmax decreased by a factor of ca. 11. The c-A myosin S1 complex provides the hitherto smallest model of actomyosin, which appears promising for crystallization and X-ray analysis.  相似文献   

14.
Current theory and experiments describing actin polymerization suggest that site-specific cleavage of bound nucleotide following F-actin filament formation causes the barbed ends of microfilaments to be capped first with ATP subunits, then with ADP bound to inorganic phosphate (ADP.Pi) at steady-state. The barbed ends of depolymerizing filaments consist of ADP subunits. The decrease in stability of the barbed-end cap accompanying the transition from ADP.Pi to ADP allows nucleotide hydrolysis and subsequent loss of Pi to regulate F-actin filament dynamics. We describe a novel computational model of nucleotide capping that simulates both the spatial and temporal properties of actin polymerization. This model has been used to test the effects of high filament concentration on the behavior of the ATP hydrolysis cycle observed during polymerization. The model predicts that under conditions of high microfilament concentration an ADP cap can appear during steady-state at the barbed ends of filaments. We show that the presence of the cap can be accounted for by a kinetic model and predict the relationship between the nucleotide concentration ratio [ATP]/[ADP], the F-actin filament concentration, and the steady-state distribution of barbed-end ADP cap lengths. The possible consequences of this previously unreported phenomenon as a regulator of cytoskeletal behavior are discussed.  相似文献   

15.
Cross-linking of actin filaments by filamin by means of frequency-dependent rheology yields an increase in the filament's elasticity and stiffness. Higher cross-linker (filamin) ratios are required for mean actin-filament lengths of 5-6 microm than for random-length distribution of actin filaments. The loss modulus (i.e. the viscous portion) in the region of the internal-chain dynamics [G"(omega) approximately omega(alpha)] is influenced by the cross-linking of filaments, and with an increasing molar ratio of filamin/actin a reduction of alpha is observed. Rheological measurements reveal that actin networks are already formed at the polymerizing stage at a molar ratio of filamin/actin of less than 1:100, and electron micrographs show phase separation of actin/filament networks of low density and of actin/filament bundles.  相似文献   

16.
The packing of the G-actin monomers within crystalline actin tubes was investigated at atomic detail. To achieve this, we have chosen an integrated structural approach which combines intermediate resolution electron microscopy based 3-D reconstruction and surface metal shadowing of crystalline actin tubes with atomic resolution X-ray data of the G-actin monomer. Distinct from the parallel, half-staggered packing of the actin subunits within F-actin filaments, the arrangement of actin monomers within the crystalline tubes involves antiparallel packing into dimers with p2 symmetry. Within the crystalline tubes, the actin monomers are oriented so that the filament axis runs parallel with the sheet plane and the intersubunit contacts in this direction are similar to those existing along the two long-pitch helical strands of the F-actin filament. The other intersubunit contacts within the crystalline tubes are not found in the actin filament. The ability of actin to form a variety of polymorphic oligomers is still not fully understood, and the functional implications of this variability have yet to be deciphered. Regularly packed actin assemblies such as sheets, tubes or ribbons may ultimately yield structural relationships to in vivo relevant actin oligomers such as, for example, the "lower dimer".  相似文献   

17.
We have studied the effect of gelsolin, a Ca-dependent actin-binding protein, on the microsecond rotational dynamics of actin filaments, using time-resolved phosphorescence (TPA) and absorption anisotropy (TAA) of erythrosin iodoacetamide attached to Cys374 on actin. Polymerization of actin in the presence of gelsolin resulted in substantial increases in the rate and amplitude of anisotropy decay, indicating increased rotational motion. Analysis indicates that the effect of gelsolin cannot be explained by increased rates of overall (rigid-body) rotations of shortened filaments, but reflects changes in intra-filament structure and dynamics. We conclude that gelsolin induces (1) a 10 degrees change in the orientation of the absorption dipole of the probe relative to the actin filament, indicating a conformational change in actin, and (2) a threefold decrease in torsional rigidity of the filament. This result, which is consistent with complementary electron microscopic observations on the same preparations, directly demonstrates long-range cooperativity in F-actin, where a conformational change induced by the binding of a single gelsolin molecule to the barbed end is propagated along inter-monomer bonds throughout the actin filament.  相似文献   

18.
The RHO1 gene encodes a homolog of mammalian RhoA small G protein in the yeast Saccharomyces cerevisiae. We have shown that Bni1p is one of the downstream targets of Rho1p and regulates reorganization of the actin cytoskeleton through the interaction with profilin, an actin monomer-binding protein. A Bni1p-binding protein was affinity purified from the yeast cytosol fraction and was identified to be Tef1p/Tef2p, translation elongation factor 1alpha (EF1alpha). EF1alpha is an essential component of the protein synthetic machinery and also possesses the actin filament (F-actin)-binding and -bundling activities. EF1alpha bound to the 186 amino acids region of Bni1p, located between the FH1 domain, the proline-rich profilin-binding domain, and the FH2 domain, of which function is not known. The binding of Bni1p to EF1alpha inhibited its F-actin-binding and -bundling activities. The BNI1 gene deleted in the EF1alpha-binding region did not suppress the bni1 bnr1 mutation in which the actin organization was impaired. These results suggest that the Rho1p-Bni1p system regulates reorganization of the actin cytoskeleton through the interaction with both EF1alpha and profilin.  相似文献   

19.
In the Drosophila flight muscle actin mutant E93K there is a charge reversal on the surface of actin close to the proposed position of tropomyosin when it is in the off state. Using a quantitative in vitro motility assay we have found that the wild type Drosophila ACT88F actin behaved like rabbit skeletal muscle actin when tropomyosin and troponin were added at pCa5 and pCa9. In contrast the effect of tropomyosin upon the E93K mutant actin filament movement was completely different from wild type and resembled the response of wild type with tropomyosin+troponin at pCa9 (i.e. the filaments were switched off). Velocity of E93K actin did not increase, and the fraction of filaments motile was reduced to less than 15% by adding up to 30 nM tropomyosin. When myosin subfragment-1 modified by N-ethylmaleimide was mixed with mutant E93K actin-tropomyosin filaments we observed that it restored motility of the filaments to the level observed with E93K actin alone. We conclude that electrostatic charge on the surface of domain 2 of actin plays a critical role in determining the state of actin-tropomyosin that is a central feature of the steric blocking mechanism of actin filament regulation.  相似文献   

20.
Here we describe the identification of a novel 37-kD actin monomer binding protein in budding yeast. This protein, which we named twinfilin, is composed of two cofilin-like regions. In our sequence database searches we also identified human, mouse, and Caenorhabditis elegans homologues of yeast twinfilin, suggesting that twinfilins form an evolutionarily conserved family of actin-binding proteins. Purified recombinant twinfilin prevents actin filament assembly by forming a 1:1 complex with actin monomers, and inhibits the nucleotide exchange reaction of actin monomers. Despite the sequence homology with the actin filament depolymerizing cofilin/actin-depolymerizing factor (ADF) proteins, our data suggests that twinfilin does not induce actin filament depolymerization. In yeast cells, a green fluorescent protein (GFP)-twinfilin fusion protein localizes primarily to cytoplasm, but also to cortical actin patches. Overexpression of the twinfilin gene (TWF1) results in depolarization of the cortical actin patches. A twf1 null mutation appears to result in increased assembly of cortical actin structures and is synthetically lethal with the yeast cofilin mutant cof1-22, shown previously to cause pronounced reduction in turnover of cortical actin filaments. Taken together, these results demonstrate that twinfilin is a novel, highly conserved actin monomer-sequestering protein involved in regulation of the cortical actin cytoskeleton.  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