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.
We have shown previously (Schwartz, M. A., and E. J. Luna. 1986. J. Cell Biol. 102: 2067-2075) that actin binds with positive cooperativity to plasma membranes from Dictyostelium discoideum. Actin is polymerized at the membrane surface even at concentrations well below the critical concentration for polymerization in solution. Low salt buffer that blocks actin polymerization in solution also prevents actin binding to membranes. To further explore the relationship between actin polymerization and binding to membranes, we prepared four chemically modified actins that appear to be incapable of polymerizing in solution. Three of these derivatives also lost their ability to bind to membranes. The fourth derivative (EF actin), in which histidine-40 is labeled with ethoxyformic anhydride, binds to membranes with reduced affinity. Binding curves exhibit positive cooperativity, and cross-linking experiments show that membrane-bound actin is multimeric. Thus, binding and polymerization are tightly coupled, and the ability of these membranes to polymerize actin is dramatically demonstrated. EF actin coassembles weakly with untreated actin in solution, but coassembles well on membranes. Binding by untreated actin and EF actin are mutually competitive, indicating that they bind to the same membrane sites. Hill plots indicate that an actin trimer is the minimum assembly state required for tight binding to membranes. The best explanation for our data is a model in which actin oligomers assemble by binding to clustered membrane sites with successive monomers on one side of the actin filament bound to the membrane. Individual binding affinities are expected to be low, but the overall actin-membrane avidity is high, due to multivalency. Our results imply that extracellular factors that cluster membrane proteins may create sites for the formation of actin nuclei and thus trigger actin polymerization in the cell.  相似文献   

2.
In previous equilibrium binding studies, Dictyostelium discoideum plasma membranes have been shown to bind actin and to recruit actin into filaments at the membrane surface. However, little is known about the kinetic pathway(s) through which actin assembles at these, or other, membranes. We have used actin fluorescently labeled with N-(1- pyrenyl)iodoacetamide to examine the kinetics of actin assembly in the presence of D. discoideum plasma membranes. We find that these membranes increase the rate of actin polymerization. The rate of membrane-mediated actin polymerization is linearly dependent on membrane protein concentrations up to 20 micrograms/ml. Nucleation (the association of activated actin monomers into oligomers) appears to be the primary step of polymerization that is accelerated. A sole effect on the initial salt-induced actin conformational change (activation) is ruled out because membranes accelerate the polymerization of pre- activated actin as well as actin activated in the presence of membranes. Elongation of preexisting filaments also is not the major step of polymerization facilitated by membranes since membranes stripped of all peripheral components, including actin, increase the rate of actin assembly to about the same extent as do membranes containing small amounts of endogenous actin. Acceleration of the nucleation step by membranes also is supported by an analysis of the dependence of polymerization lag time on actin concentration. The barbed ends of membrane-induced actin nuclei are not obstructed by the membranes because the barbed end blocking agent, cytochalasin D, reduces the rate of membrane-mediated actin nucleation. Similarly, the pointed ends of the nuclei are not blocked by membranes since the depolymerization rate of gelsolin-capped actin is unchanged in the presence of membranes. These results are consistent with previous observations of lateral interactions between membranes and actin filaments. These results also are consistent with two predictions from a model based on equilibrium binding studies; i.e., that plasma membranes should nucleate actin assembly and that membrane-bound actin nuclei should have both ends free (Schwartz, M. A., and E. J. Luna. 1988. J. Cell Biol. 107:201-209). Integral membrane proteins mediate the actin nucleation activity because activity is eliminated by heat denaturation, treatment with reducing agents, or proteolysis of membranes. Activity also is abolished by solubilization with octylglucoside but is reconstituted upon removal or dilution of the detergent. Ponticulin, the major actin-binding protein in plasma membranes, appears to be necessary for nucleation activity since activity is not reconstituted from detergent extracts depleted of ponticulin.  相似文献   

3.
《The Journal of cell biology》1989,109(6):2833-2840
We have investigated the association of actin with membranes isolated from rat liver. A plasma membrane-enriched fraction prepared by homogenization in a low salt/CaCl2 buffer was found to contain a substantial amount of residual actin which could be removed by treatment with 1 M Na2CO3/NaHCO3, pH 10.5. Using a sedimentation binding assay that uses gelsolin to shorten actin filaments and render membrane binding saturable (Schwartz, M. A., and E. J. Luna. 1986. J. Cell Biol. 102:2067-2075), we found that membranes stripped of endogenous actin bound 125I-actin in a specific and saturable manner. Scatchard plots of binding data were linear, indicating a single class of binding sites with a Kd of 1.6 microns; 66 micrograms actin bound/mg membrane protein at saturation. Binding of actin to liver cell membranes was negligible with unstripped membranes, was competed by excess unlabeled actin, and was greatly reduced by preheating or proteolytic digestion of the membranes. Kinetic measurements showed that binding had an initial lag phase and was strongly temperature dependent. The binding of actin to liver cell membranes was also found to be competitively inhibited by ATP and other nucleotides, including the nonhydrolyzable analogue AMP-PNP. We conclude that we have reconstituted an interaction between actin and integral membrane proteins from the rat liver. This interaction exhibits a number of distinctive features which have not been observed in other actin- membrane systems.  相似文献   

4.
H E Harris 《Biochemistry》1985,24(23):6613-6618
Actin and plasma gelsolin were covalently cross-linked with the zero-length cross-linker 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide. Two major intermolecularly linked products were identified on polyacrylamide gels. By use of 14C-labeled actin and 125I-labeled gelsolin, these were shown to be the 1:1 and 2:1 complexes of actin with gelsolin, respectively. The higher molecular weight complex predominated under all conditions tested including the presence and absence of Ca2+. In titration experiments in which actin at different concentrations was reacted with a fixed concentration of gelsolin, end points were obtained for the formation of both cross-linked species at about two actins per gelsolin, implying that a 2:1 noncovalent complex is cross-linked. In 0.1 mM Ca2+, the extent of cross-linking was independent of protein concentration down to 50 nM gelsolin. At low Ca2+ concentrations (less than 10(-8)M), the extent of cross-linking was very much reduced at micromolar gelsolin and fell to zero at about 100 nM gelsolin. The binding of actin to gelsolin to give a cross-linkable complex is therefore very strong at 0.1 mM Ca2+ but much weaker at low Ca2+ concentrations.  相似文献   

5.
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   

6.
We have characterized the interaction of bovine pancreatic deoxyribonuclease I (DNase I) with the filamentous (F-)actin of red cell membrane skeletons stabilized with phalloidin. The hydrolysis of [3H]DNA was used to assay DNase I. We found that DNase I bound to a homogenous class of approximately equal to 2.4 X 10(4) sites/skeleton with an association rate constant of approximately 1 X 10(6) M-1 S-1 and a KD of 1.9 X 10(-9) M at 20 degrees C. Phalloidin lowered the dissociation constant by approximately 1 order of magnitude. The DNase I which sedimented with the skeletons was catalytically inactive but could be reactivated by dissociation from the actin. Actin and DNA bound to DNase I in a mutually exclusive fashion without formation of a ternary complex. Phalloidin-treated red cell F-actin resembled rabbit muscle G-actin in all respects tested. Since the DNase I binding capacity of the skeletons corresponded to the number of actin protofilaments previously estimated by other methods, it seemed likely that the enzyme binding site was confined to one end of the filament. We confirmed this premise by showing that elongating the red cell filaments with rabbit muscle actin monomers did not appreciably add to their capacity to bind or inhibit DNase I. Saturation of skeletons with cytochalasin D or gelsolin, avid ligands for the barbed end of actin filaments, did not reduce their binding of DNase I. Furthermore, neither cytochalasin D nor DNase I alone blocked all of the sites for addition of monomeric pyrene-labeled rabbit muscle G-actin to phalloidin-treated skeletons; however, a combination of the two agents did so. In the presence of phalloidin, the polymerization of 300 nM pyrenyl actin on nuclei constructed from 5 nM gelsolin and 25 nM rabbit muscle G-actin was completely inhibited by 35 nM DNase I but not by 35 nM cytochalasin D. We conclude that DNase I associates uniquely with and caps the pointed (slow-growing or negative) end of F-actin. These results imply that the amino-terminal, DNase I-binding domain of the actin protomer is oriented toward the pointed end and is buried along the length of the actin filament.  相似文献   

7.
We prepared a probe of radiolabeled, glutaraldehyde cross-linked filamentous actin (F-actin) to study binding of actin to membranes of Dictyostelium discoideum. The probe bound to membranes or detergent extracts of membranes with a high affinity and in a saturable manner. The binding could be reduced by boiling of either the actin probe or the membranes, or by addition of excess native F-actin, but not by addition of an equivalent amount of bovine serum albumin, to the assay. The probe labeled several proteins when used to overlay sodium dodecyl sulfate gels of Dictyostelium membranes. One of these labeled proteins was a 24,000-mol-wt protein (p24), which was soluble only in the presence of a high concentration of sodium deoxycholate (5%, wt/vol) at room temperature or above. The p24 was purified by selective detergent extraction and column chromatography. When tested in a novel two-phase binding assay, p24 bound both native monomeric actin (G-actin) and F-actin in a specific manner. In this assay, G-actin bound p24 with a submicromolar affinity.  相似文献   

8.
Gelsolin and calponin are cytoskeletal and signalling proteins that form a tight 1:1 complex (GCC). We show that calponin within the GCC inhibits the rate of gelsolin mediated nucleation of actin polymerization. The actin-binding function of calponin is ablated within the GCC as the actin-binding site overlaps with one of the gelsolin binding sites. The structure of filaments that result from nucleation by GCC are different to those nucleated by gelsolin alone in that they are longer, loosely bundled and stain heterogeneously with phalloidin. GCC nucleated filaments appear contorted and wrap around each to form the loose bundles.  相似文献   

9.
Actin cleaved by the protease from Escherichia coli A2 strain between Gly42 and Val43 (ECP-actin) is no longer polymerizable when it contains Ca2+ as a tightly bound cation, but polymerizes when Mg2+ is bound. We have investigated the interactions of gelsolin with this actin with regard to conformational changes in the actin molecule induced by the binding of gelsolin. ECP-(Ca)actin interacts with gelsolin in a manner similar to that in which it reacts with intact actin, and forms a stoichiometric 2:1 complex. Despite the nonpolymerizability of ECP-(Ca)actin, this complex can act as a nucleus for the polymerization of intact actin, thus indicating that upon interaction with gelsolin, ECP-(Ca)actin undergoes a conformational change that enables its interaction with another actin monomer. By gel filtration and fluorometry it was shown that the binding of at least one of the ECP-cleaved actins to gelsolin is considerably weaker than of intact actin, suggesting that conformational changes in subdomain 2 of actin monomer may directly or allosterically affect actin-gelsolin interactions. On the other hand, interaction with gelsolin changes the conformation of actin within the DNase I-binding loop, as indicated by inhibition of limited proteolysis of actin by ECP and subtilisin. Cross-linking experiments with gelsolin-nucleated actin filaments using N,N-phenylene-bismaleimide (which cross-links adjacent actin monomers between Cys374 and Lys191) reveal that gelsolin causes a significant increase in the yield of the 115-kDa cross-linking product, confirming the evidence that gelsolin stabilizes or changes the conformation of the C-terminal region of the actin molecule, and these changes are propagated from the capped end along the filament. These results allow us to conclude that nucleation of actin polymerization by gelsolin is promoted by conformational changes within subdomain 2 and at the C-terminus of the actin monomer.  相似文献   

10.
The acrosomal reaction (AR) is a regulated sperm exocytotic process that involves fusion of the plasma membrane (PM) with the outer acrosomal membrane (OAM). Our group has described F-actin cytoskeletons associated to these membranes. It has been proposed that in regulated exocytosis, a cortical cytoskeleton acts as a barrier that obstructs membrane fusion, and must be disassembled for exocytosis to occur. Actin-severing proteins from the gelsolin family have been considered to break this barrier. The present study attempted to determine if gelsolin has a function in guinea pig sperm capacitation and AR. By indirect immunofluorescence (IIF), gelsolin was detected in the apical and postacrosomal regions of the head and in the flagellum in both capacitated and non-capacitated guinea pig spermatozoa. By Western blotting, gelsolin was detected in isolated PM and OAM of non-capacitated spermatozoa. Gelsolin and actin were detected in a mixture of PM-OAM obtained by sonication, and both proteins were absent in membranes of capacitated spermatozoa. Inhibition of three different pathways of PIP2 hydrolysis during capacitation did not cancel gelsolin loss from membranes. Gelsolin was detected by Western blotting associated to membrane cytoskeletons obtained after phalloidin F-actin stabilization and Triton-X treatment; additionally, by immunoprecipitation, it was shown that gelsolin is associated with actin. By electron microscopy we observed that skeletons disassemble during capacitation, but phalloidin prevents disassembly. A three-dimensional skeleton was observed that apparently joins PM with OAM. Exogenous gelsolin stimulates AR assayed in a permeabilized spermatozoa model. Results suggest that gelsolin disassembles F-actin cytoskeletons during capacitation, promoting AR.  相似文献   

11.
We describe the interaction of pure brain tubulin with purified membranes specialized in different cell functions, i.e., plasma membranes and mitochondrial membranes from liver and secretory granule membranes from adrenal medulla. We studied the tubulin-binding activity of cellular membranes using a radiolabeled ligand-receptor assay and an antibody retention assay. The tubulin-membrane interaction was time- and temperature-dependent, reversible, specific, and saturable. The binding of tubulin to membranes appears to be specific since acidic proteins such as serum albumin or actin did not interfere in the binding process. The apparent overall affinity constant of the tubulin- membrane interaction ranged between 1.5 and 3.0 X 10(7) M-1; similar values were obtained for the three types of membranes. Tubulin bound to membranes was not entrapped into vesicles since it reacted quantitatively with antitubulin antibodies. At saturation of the tubulin-binding sites, the amount of reversibly bound tubulin represents 5-10% by weight of membrane protein (0.4-0.9 nmol tubulin/mg membrane protein). The high tubulin-binding capacity of membranes seems to be inconsistent with a 1:1 stoichiometry between tubulin and a membrane component but could be relevant to a kind of tubulin assembly. Indeed, tubulin-membrane interaction had some properties in common with microtubule formation: (a) the association of tubulin to membranes increased with the temperature, whereas the dissociation of tubulin- membrane complexes increased by decreasing temperature; (b) the binding of tubulin to membranes was prevented by phosphate buffer. However, the tubulin-membrane interaction differed from tubulin polymerization in several aspects: (a) it occurred at concentrations far below the critical concentration for polymerization; (b) it was not inhibited at low ionic strength and (c) it was colchicine-insensitive. Plasma membranes, mitochondrial membranes, and secretory granule membranes contained tubulin as an integral component. This was demonstrated on intact membrane and on Nonidet P-40 solubilized membrane protein using antitubulin antibodies in antibody retention and radioimmune assays. Membrane tubulin content varied from 2.2 to 4.4 micrograms/mg protein. The involvement of membrane tubulin in tubulin-membrane interactions remains questionable since erythrocyte membranes devoid of membrane tubulin exhibited a low (one-tenth of that of rat liver plasma membranes) but significant tubulin-binding activity. These results show that membranes specialized in different cell functions possess high- affinity, large-capacity tubulin-binding sites...  相似文献   

12.
Various concentrations of gelsolin (25-100 nM) were added to 2 microM polymerized actin. The concentrations of free calcium were adjusted to 0.05-1.5 microM by EGTA/Ca2+ buffer. Following addition of gelsolin actin depolymerization was observed that was caused by dissociation of actin subunits from the pointed ends of treadmilling actin filaments and inhibition by gelsolin of polymerization at barbed ends. The time course of depolymerization revealed an initial lag phase that was followed by slow decrease of the concentration of polymeric actin to reach the final steady state polymer and monomer concentration. The initial lag phase was pronounced at low free calcium and low gelsolin concentrations. On the basis of quantitative analysis the kinetics of depolymerization could be interpreted as capping, i.e. binding of gelsolin to the barbed ends of actin filaments and subsequent inhibition of polymerization, rather than severing. The main argument for this conclusion was that even gelsolin concentrations (100 nM) that exceed the concentration of filament ends ( approximately 2 nM), cause the filaments to depolymerize at a rate that is similar to the rate of depolymerization of the concentration of pointed ends existing before addition of gelsolin. The rate of capping is directly proportional to the free calcium concentration. These experiments demonstrate that at micromolar and submicromolar free calcium concentrations gelsolin acts as a calcium-regulated capping protein but not as an actin filament severing protein, and that the calcium binding sites of gelsolin which regulate the various functions of gelsolin (capping, severing and monomer binding), differ in their calcium affinity.  相似文献   

13.
Stable oligomers of filamentous actin were obtained by cross-linking F-actin with 1,4-N,N'-phenylenedimaleimide and depolymerization with excess segment-1 of gelsolin. Segment-1-bound and cross-linked actin oligomers containing either two or three actin subunits were purified and shown to nucleate actin assembly. Kinetic assembly data from mixtures of monomeric actin and the actin oligomers fit a nucleation model where cross-linked actin dimer or trimer reacts with an actin monomer to produce a competent nucleus for filament assembly. We report the three-dimensional structure of the segment-1-actin hexamer containing three actin subunits, each with a tightly bound ATP. Comparative analysis of this structure with twelve other actin structures provides an atomic level explanation for the preferential binding of ATP by the segment-1-complexed actin. Although the structure of segment-1-bound actin trimer is topologically similar to the helical model of F-actin (1), it has a distorted symmetry compared with that of the helical model. This distortion results from intercalation of segment-1 between actin protomers that increase the rise per subunit and rotate each of the actin subunits relative to their positions in F-actin. We also show that segment-1 of gelsolin is able to sever actin filaments, although the severing activity of segment-1 is significantly lower than full-length gelsolin.  相似文献   

14.
The importance of actin hydrophobic loop 262-274 dynamics to actin polymerization and filament stability has been shown recently with the use of the yeast mutant actin L180C/L269C/C374A, in which the hydrophobic loop could be locked in a “parked” conformation by a disulfide bond between C180 and C269. Such a cross-linked globular actin monomer does not form filaments, suggesting nucleation and/or elongation inhibition. To determine the role of loop dynamics in filament nucleation and/or elongation, we studied the polymerization of the cross-linked actin in the presence of cofilin, to assist with actin nucleation, and with phalloidin, to stabilize the elongating filament segments. We demonstrate here that together, but not individually, phalloidin and cofilin co-rescue the polymerization of cross-linked actin. The polymerization was also rescued by filament seeds added together with phalloidin but not with cofilin. Thus, loop immobilization via cross-linking inhibits both filament nucleation and elongation. Nevertheless, the conformational changes needed to catalyze ATP hydrolysis by actin occur in the cross-linked actin. When actin filaments are fully decorated by cofilin, the helical twist of filamentous actin (F-actin) changes by ∼ 5° per subunit. Electron microscopic analysis of filaments rescued by cofilin and phalloidin revealed a dense contact between opposite strands in F-actin and a change of twist by ∼ 1° per subunit, indicating either partial or disordered attachment of cofilin to F-actin and/or competition between cofilin and phalloidin to alter F-actin symmetry. Our findings show an importance of the hydrophobic loop conformational dynamics in both actin nucleation and elongation and reveal that the inhibition of these two steps in the cross-linked actin can be relieved by appropriate factors.  相似文献   

15.
Orbán J  Lorinczy D  Hild G  Nyitrai M 《Biochemistry》2008,47(15):4530-4534
Actin plays important roles in eukaryotic cell motility. During actin polymerization, the actin-bound ATP is hydrolyzed to ADP and P i. We carried out differential scanning calorimetry experiments to characterize the cooperativity of the stabilizing effect of phalloidin on actin filaments in their ADP.P i state. The ADP.P i state was mimicked by using ADP.BeF x or ADP.AlF 4. The results showed that the binding of the nucleotide analogues or phalloidin stabilized the actin filaments to a similar extent when added separately. Phalloidin binding to ADP.BeF x- or ADP.AlF 4-actin filaments further stabilized them, indicating that the mechanism by which phalloidin and the nucleotide analogues affect the filament structure was different. The results also showed that the stabilization effect of phalloidin binding to ADP.BeF x or ADP.AlF 4-bound actin filaments was not cooperative. Since the effect of phalloidin binding was cooperative in the absence of these nucleotide analogues, these results suggest that the binding of ADP.BeF x or ADP.AlF 4 to the actin modified the protomer-protomer interactions along the actin filaments.  相似文献   

16.
The rate of association of actin with gelsolin was measured at various Ca2+ and ATP concentrations. The fraction of Ca2+-activated gelsolin was determined by quantitative evaluation of the association rates thereby assuming that Ca2+-binding gelsolin associates with actin and Ca2+-free gelsolin does not. A plot of the fraction of Ca2+-activated gelsolin vs. the free Ca2+ concentration revealed a sigmoidal shape suggesting that co-operative binding of Ca2+ ions is required for activation of gelsolin. A good fit of the experimental data by calculated binding curves was obtained if two Ca2+ ions were assumed to bind to actin in a highly co-operative manner. ATP decreased the rate of association of gelsolin with actin and bound to gelsolin at a low affinity (Kd = 32 microm for Ca2+-free and Kd = 400 microm for Ca2+-activated gelsolin). In contrast, a 1 : 1 gelsolin-actin complex was found to be activated for association with actin by a single Ca2+ ion in a non-co-operative manner.  相似文献   

17.
The beta-thymosins are intracellular monomeric (G-)actin sequestering proteins forming 1:1 complexes with G-actin. Here, we analysed the interaction of thymosin beta(4) with F-actin. Thymosin beta(4) at 200 microM was chemically cross-linked to F-actin. In the presence of phalloidin, the chemically cross-linked actin:thymosin beta(4) complex was incorporated into F-actin. These mixed filaments were of normal appearance when inspected by conventional transmission electron microscopy after negative staining. We purified the chemically cross-linked actin:thymosin beta(4) complex, which polymerised only when phalloidin and the gelsolin:2-actin complex were present simultaneously. Using scanning transmission electron microscopy, the mass-per-length of control and actin:thymosin beta(4) filaments was found to be 16.0(+/-0.8) kDa/nm and 18.0(+/-0.9) kDa/nm, respectively, indicating an increase in subunit mass of 5.4 kDa. Analysis of the helical parameters revealed an increase of the crossover spacing of the two right-handed long-pitch helical strands from 36.0 to 40.5 nm. Difference map analysis of 3-D helical reconstruction of control and actin:thymosin beta(4) filaments yielded an elongated extra mass. Qualitatively, the overall size and shape of the difference mass were compatible with published data of the atomic structure of thymosin beta(4). The deduced binding sites of thymosin beta(4) to actin were in agreement with those identified previously. However, parts of the difference map might represent subtle conformational changes of both proteins occurring upon complex formation.  相似文献   

18.
Differential scanning calorimetry was used to examine the effects of cofilin on the thermal unfolding of actin. Stoichiometric binding increases the thermal stability of both G- and F-actin but at sub-saturating concentrations cofilin destabilizes F-actin. At actin:cofilin molar ratios of 1.5-6 the peaks corresponding to stabilized (66-67 degrees C) and destabilized (56-57 degrees C) F-actin are observed simultaneously in the same thermogram. Destabilizing effects of sub-saturating cofilin are highly cooperative and are observed at actin:cofilin molar ratios as low as 100:1. These effects are abolished by the addition of phalloidin or aluminum fluoride. Conversely, at saturating concentrations, cofilin prevents the stabilizing effects of phalloidin and aluminum fluoride on the F-actin thermal unfolding. These results suggest that cofilin stabilizes those actin subunits to which it directly binds, but destabilizes F-actin with a high cooperativity in neighboring cofilin-free regions.  相似文献   

19.
Plasmodium falciparum, the etiologic agent of malaria, is a facultative intracellular parasite of the phylum Apicomplexa. A limited turnover of microfilaments takes place beneath the parasite plasma membrane, but the cytoplasm of apicomplexans is virtually devoid of F-actin. We produced Plasmodium actin in yeast. Purified recombinant Plasmodium actin polymerized inefficiently unless both gelsolin and phalloidin were added. The resulting actin polymers appeared fragmented in the fluorescence microscope. Plasmodium actin bound DNaseI about 200 times weaker than bovine non-muscle actin. Our findings suggest that the unique properties of Plasmodium actin can explain some of the unusual features of apicomplexan parasite microfilaments.  相似文献   

20.
AimsFormation of different protrusive structures by migrating cells is driven by actin polymerization at the plasma membrane region. Gelsolin is an actin binding protein controlling the length of actin filaments by its severing and capping activity. The main goal of this study was to determine the effect of gelsolin expression on the migration of human colon adenocarcinoma LS180 and melanoma A375 cells.Main methodsColon adenocarcinoma cell line LS180 was stably transfected with plasmid containing human cytoplasmic gelsolin cDNA tagged to enhanced green fluorescence protein (EGFP). Melanoma A375 cells were transfected with siRNAs directed against gelsolin. Real-time PCR and Western blotting were used to determine the level of gelsolin. The ability of actin to inhibit DNase I activity was used to quantify monomeric and total actin level and calculate the state of actin polymerization. Fluorescence confocal microscopy was applied to observe gelsolin and vinculin distribution along with actin cytoskeleton organization.Key findingsIncreased level of gelsolin expression leads to its accumulation at the submembranous region of the cell accompanied by distinct changes in the state of actin polymerization and an increase in the migration of LS180 cells. In addition, LS180 cells overexpressing gelsolin form podosome-like structures as indicated by vinculin redistribution and its colocalization with gelsolin and actin. Downregulation of gelsolin expression in melanoma A375 cells significantly reduces their migratory potential.SignificanceOur experimental data indicate that alterations in the expression level of gelsolin and its subcellular distribution may be directly responsible for determining migration capacity of human cancer cells.  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