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.
The status of platelet actin has been studied analytically by the DNase-I inhibition assay. Exposure of platelets to diamide, which oxidizes sulphydryl groups, results in an increase of filamentous actin. Neutralization of the effect of diamide by addition of 2-mercaptopropionylglycine (2-MPG) or washing out the excess of diamide is associated with a normalization of the cellular actin status. These findings strongly suggest that the redox state of platelets is somehow involved in the process of actin polymerization. Alterations of the redox state in metabolically altered platelets could therefore account for changes in the G- to F-actin equilibrium. In the thrombin-induced actin polymerization the redox state of the cell seems to be not involved, since 2-MPG preincubation of platelets (1.25 mM, 30 min) does not inhibit the filament assembly.  相似文献   

2.
The production of a soluble actomyosin complex would be a significant step toward elucidating molecular interactions responsible for biological movement. We took a systematic approach to producing soluble long-pitch actin dimers that are polymerization-deficient yet retain biological protein-protein interactions, including myosin binding. Actin mutant proteins and chemical crosslinking were combined with different polymerization inhibition strategies, including ADP-ribosylation, or the use of a polymerization-deficient actin mutant protein. While all of the long-pitch actin dimers retained interactions reflective of F-actin activity, each displayed different interactions with myosin. Myosin did not interact productively with long-pitch actin dimers capped with DNase-I, and led to filament formation of unmodified long-pitch actin dimers or dimers possessing a polymerization-deficient actin subunit. However, ADP-ribosylated long-pitch actin dimers interacted with myosin, giving this dimer great potential for producing a soluble actomyosin complex, which could greatly improve our understanding of the molecular basis of movement in cells, tissues, and organisms.  相似文献   

3.
Several lines of evidence indicate that sucrose synthase (SuSy) binds both G- and F-actin: (i) presence of SuSy in the Triton X-100-insoluble fraction of microsomal membranes (i.e. crude cytoskeleton fraction); (ii) co-immunoprecipitation of actin with anti-SuSy monoclonal antibodies; (iii) association of SuSy with in situ phalloidin-stabilized F-actin filaments; and (iv) direct binding to F-actin, polymerized in vitro. Aldolase, well known to interact with F-actin, interfered with binding of SuSy, suggesting that a common or overlapping binding site may be involved. We postulate that some of the soluble SuSy in the cytosol may be associated with the actin cytoskeleton in vivo.  相似文献   

4.
Protein kinase C alpha (PKC alpha) participates in F-actin remodeling during phagocytosis and phagosomal maturation in macrophages. Leishmania donovani promastigotes, which inhibit phagosomal maturation, cause accumulation of periphagosomal F-actin instead of the disassembly observed around other prey [Cell. Microbiol. 7 (2001) 439]. This accumulation is induced by promastigote lipophosphoglycan (LPG), which has several effects on macrophages including inhibition of PKC alpha. To investigate a possible connection between PKC alpha and LPG's effects on actin dynamics, we utilized RAW264.7 macrophages overexpressing dominant-negative PCK alpha (DN PKC alpha). We found increased cortical F-actin and decreased phagocytic capacity, as well as defective periphagosomal F-actin breakdown and inhibited phagosomal maturation in the DN PKC alpha-overexpressing cells, effects similar to those seen in controls subjected to LPG-coated prey. The results indicate that PKC alpha is involved in F-actin turnover in macrophages and that PKC alpha-dependent breakdown of periphagosomal F-actin is required for phagosomal maturation, and endorse the hypothesis that intracellular survival of L. donovani involves inhibition of PKC alpha by LPG.  相似文献   

5.
We have examined the role of endogenous 70-kDa S6 kinase (p70(S6K)) in actin cytoskeletal organization and cell migration in Swiss 3T3 fibroblasts. Association of p70(S6K) with the actin cytoskeleton was demonstrated by cosedimentation of p70(S6K) with F-actin and by subcellular fractionation in which p70(S6K) activity was measured in the F-actin cytoskeletal fraction. Immunocytochemical studies showed that p70(S6K), Akt1, PDK1, and p85 phosphoinositide 3-kinase (PI 3-kinase) were localized to the actin arc, a caveolin-enriched cytoskeletal structure located at the leading edge of migrating cells. Using a phospho-specific antibody to mammalian target of rapamycin (mTOR), we find that activated mTOR is enriched at the actin arc, suggesting that activation of the p70(S6K) signaling pathway is important to cell migration. Using the actin arc to assess migration, epidermal growth factor (EGF) stimulation was found to induce actin arc formation, an effect that was blocked by rapamycin treatment. We show further that actin stress fibers may function to down-regulate p70(S6K). Fibronectin stimulated stress fiber formation in the absence of growth factors and caused an inactivation of p70(S6K). Conversely, cytochalasin D and the Rho kinase inhibitor Y-27632, both of which cause stress fiber disruption, increased p70(S6K) activity. These studies provide evidence that the p70(S6K) pathway is important for signaling at two F-actin microdomains in cells and regulates cell migration.  相似文献   

6.
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   

7.
Cofilin, a key regulator of actin filament dynamics, binds to G- and F-actin and promotes actin filament turnover by stimulating depolymerization and severance of actin filaments. In this study, cytochalasin D (CytoD), a widely used inhibitor of actin dynamics, was found to act as an inhibitor of the G-actin-cofilin interaction by binding to G-actin. CytoD also inhibited the binding of cofilin to F-actin and decreased the rate of both actin polymerization and depolymerization in living cells. CytoD altered cellular F-actin organization but did not induce net actin polymerization or depolymerization. These results suggest that CytoD inhibits actin filament dynamics in cells via multiple mechanisms, including the well-known barbed-end capping mechanism and as shown in this study, the inhibition of G- and F-actin binding to cofilin.  相似文献   

8.
A Lees  J G Haddad  S Lin 《Biochemistry》1984,23(13):3038-3047
Actin depolymerizing activity in serum can be attributed to the two proteins brevin and vitamin D binding protein (DBP). To investigate their mechanisms of action, we used a number of techniques, including procedures involving the fluorescent pyrene-labeled actin probe, to compare the interaction of the two proteins with G- and F-actin in vitro. With a fluorescence enhancement assay, we determined that brevin forms a 1:2 complex and DBP forms a 1:1 complex with pyrene-G-actin. We also found that both proteins reduce the viscosity of F-actin measured with high-shear and low-shear viscometers, with brevin effective at much lower concentrations than DBP. In polymerization experiments, brevin inhibits filament elongation at substoichiometric levels by inhibiting monomer addition at the barbed end but can also accelerate polymerization by nucleating assembly of filaments which grow from the pointed end. DBP does not nucleate filament assembly and inhibits filament elongation at either end only at near-stoichiometric levels. Brevin, but not DBP, accelerates disassembly of filaments diluted into a depolymerizing medium. This is consistent with the capability of brevin to sever preformed filaments associated with erythrocyte membranes and to increase the number of filament ends as estimated by a cytochalasin binding assay. In steady-state experiments involving the use of pyrene-actin, brevin produces only a small increase in the apparent monomer concentration when the critical concentrations at the two ends of the filaments are the same (i.e., in 0.1 M KCl). However, when the critical concentration at the pointed end is higher than that at the barbed end (i.e., in 2 mM MgCl2), low molar ratios of brevin sharply increase the monomer concentration to the critical concentration of the pointed end. This allows substoichiometric amounts of brevin to completely depolymerize filaments when the total actin concentration is at or below that of the pointed end. In contrast to brevin, DBP increases the amount of nonfilamentous actin in a stoichiometric and dose-dependent manner regardless of the nature of the salt in the medium. We conclude from this study that brevin is similar in its mechanism of action to other proteins known to bind to the barbed end of filaments and that DBP is related in its action to proteins that complex monomers and prevent them from participating in the polymerization process.  相似文献   

9.
Mammalian spermatozoa should reside in the female reproductive tract for a certain time before gaining the ability to fertilize. During this time, the spermatozoa undergo a series of biochemical processes collectively called capacitation. We recently demonstrated that actin polymerization is a necessary step in the cascade leading to capacitation. We demonstrate here for the first time a role for phospholipase D (PLD) in the induction of actin polymerization and capacitation in spermatozoa. The involvement of PLD is supported by specific inhibition of F-actin formation during sperm capacitation by PLD inhibitors and the stimulation of fast F-actin formation by exogenous PLD or phosphatidic acid (PA). Moreover, PLD activity is enhanced during capacitation before actin polymerization. Protein kinase A (PKA), known to be active in sperm capacitation, and protein kinase C (PKC), involved in the acrosome reaction, can both activate PLD and actin polymerization. We suggest that PKA- and PKC-dependent signal transduction pathways can potentially lead to PLD activation; however, under physiological conditions, actin polymerization depends primarily on PKA activity. Activation of PKA during capacitation causes inactivation of phospholipase C, and as a result, PKC activation is prevented. It appears that PKA activation promotes sperm capacitation whereas early activation of PKC during capacitation would jeopardize this process.  相似文献   

10.
The effect of novel cytotoxic marine macrolide, amphidinolide H (Amp-H), on actin dynamics was investigated in vitro. Amp-H attenuated actin depolymerization induced by diluting F-actin. This effect remained after washing out of unbound Amp-H by filtration. In the presence of either Amp-H or phalloidin, lag phase, which is the rate-limiting step of actin polymerization, was shortened. Phalloidin decreased the polymerization-rate whereas Amp-H did not. Meanwhile, the effects of both compounds were the same when barbed end of actin was capped by cytochalasin D. Quartz crystal microbalance system revealed interaction of Amp-H with G-actin and F-actin. Amp-H also enhanced the binding of phalloidin to F-actin. We concluded that Amp-H stabilizes actin in a different manner from that of phalloidin and serves as a novel pharmacological tool for analyzing actin-mediated cell function.  相似文献   

11.
Oda T  Namba K  Maéda Y 《Biophysical journal》2005,88(4):2727-2736
Knowledge of the phalloidin binding position in F-actin and the relevant understanding of the mechanism of F-actin stabilization would help to define the structural characteristics of the F-actin filament. To determine the position of bound phalloidin experimentally, x-ray fiber diffraction data were obtained from well-oriented sols of F-actin and the phalloidin-F-actin complex. The differences in the layer-line intensity distributions, which were clearly observed even at low resolution (8 A), produced well-resolved peaks corresponding to interphalloidin vectors in the cylindrically averaged difference-Patterson map, from which the radial binding position was determined to be approximately 10 A from the filament axis. Then, the azimuthal and axial positions were determined by single isomorphous replacement phasing and a cross-Patterson map in radial projection to be approximately 84 degrees and 0.5 A relative to the actin mass center. The refined position was close to the position found by prior researchers. The position of rhodamine attached to phalloidin in the rhodamine-phalloidin-F-actin complex was also determined, in which the conjugated Leu(OH)(7) residue was found to face the outside of the filament. The position and orientation of the bound phalloidin so determined explain the increase in the interactions between long-pitch strands of F-actin and would also account for the inhibition of phosphate release, which might also contribute to the F-actin stabilization. The method of analysis developed in this study is applicable for the determination of binding positions of other drugs, such as jasplakinolide and dolastatin 11.  相似文献   

12.
Fragmin from plasmodium of Physarum polycephalum binds G-actin and severs F-actin in the presence of Ca2+ over 10(-6) M. The fragmin-actin complex consisting of fragmin and G-actin nucleates actin polymerization and caps the barbed (fast growing) end of F-actin, regardless of the concentrations of Ca2+, and the actin filaments are shortened. Actin kinase purified from plasmodium abolishes the nucleation and capping activities of the complex by phosphorylating actin of the fragmin-actin complex (Furuhashi, K., and Hatano, S. (1990) J. Cell. Biol. 111, 1081-1087). This inactivation of the complex leads to production of long actin filaments. We obtained evidence that Physarum actin is phosphorylated by actin kinase at Thr-201, and probably at Thr-202 and/or Thr-203, with 1 mol of phosphate distributed among them. This finding raises the possibility that the site of phosphorylation, Thr-201 to Thr-203, is positioned on the pointed (slow growing) end domain of the actin molecule, because growth of actin filaments from the fragmin-actin complex occurs only from the pointed end. These observations are consistent with a model of the three-dimensional structure of G-actin. Inactivation of the fragmen-actin complex may follow phosphorylation of the pointed end domain of actin.  相似文献   

13.
14.
Phosphatidylinositol 3-kinase (PI3K) activity is known to be required for the extension of embryonic sensory axons. Inhibition of PI3K has also been shown to mediate axon retraction and growth cone collapse in response to semaphorin 3A. However, the effects of inhibiting PI3K on the neuronal cytoskeleton are not well characterized. We have previously reported that semaphorin 3A-induced axon retraction involves activation of myosin II, the formation of an intra-axonal F-actin bundle cytoskeleton, and blocks the formation of F-actin patches that serve as precursors to filopodial formation in axons. We now report that inhibition of PI3K results in activation of myosin II in axons. Inhibition of myosin II activity, or its upstream regulatory kinase RhoA-kinase, blocked axon retraction induced by inhibition of PI3K. In addition, inhibition of PI3K also induced intra-axonal F-actin bundles, which likely serve as a substratum for myosin II-based force generation during axon retraction. In axons, filopodia are formed from axonal F-actin patch precursors. Analysis of axonal F-actin patch formation in eYFP-actin expressing neurons revealed that inhibition of PI3K blocked formation of axonal F-actin patches, and thus filopodial formation. These data provide insights into the regulation of the neuronal cytoskeleton by PI3K and are consistent with the notion that decreased levels of PI3K activity mediate axon retraction and growth cone collapse in response to semaphorin 3A.  相似文献   

15.
We have re-examined the Ca(++)-dependent interaction of an intestinal microvillar 95- kdalton protein (MV-95K) and actin using the isolated acrosomal process bundles from limulus sperm. Making use of the processes as nuclei for assembling actin filaments, we quantitatively and qualitatively examined MV-95K’s effect on filament assembly and on F- actin, both in the presence and in the absence of Ca(++). The acrosomal processes are particularly advantageous for this approach because they nucleate large numbers of filaments, they are extremely stable, and their morphology can be used to determine the polarity of any nucleated filaments. When filament nucleation was initiated in the presence of MV-95K and the absence of Ca(++), there was biased filament assembly from the bundle ends. The calculated elongation rates from both the barbed and pointed filament ends were virtually indistinguishable from control preparations. In the presence of Ca(++), MV-95K completely inhibited filament assembly from the barbed filament end without affecting the initial rate of assembly from the pointed filament end. The inhibition of assembly results from MV-95K binding to and capping the barbed filament end, thereby preventing monomer addition. This indicates that, while MV-95K is a potent nucleator of actin assembly, it is also a potent inhibitor of actin filament elongation. To examine the effects of MV-95K on F-actin in the presence of Ca(++), we developed an assay where MV-95K is added to filaments previously assembled from acrosomal processes without causing filament breakage during mixing. These results clearly demonstrated that rapid filament shortening by MV-95K results through a mechanism of disrupting intrafilament monomer-monomer interactions. Finally, we show that tropomyosin-containing actin filaments are insensitive to cutting, but not to capping, by MV-95K in the presence of Ca(++).  相似文献   

16.
F-Actin was digested with alpha-chymotrypsin in 6 M urea, and two peptide fragments from subdomain 4 of actin molecule [Kabsch, W., Mannherz, H.G., Suck, D., Pai, E.F., & Holmes K.C. (1990) Nature 347, 37-44] were purified by reverse-phase HPLC and Sephadex G-50 gel filtration. The peptide fragments were identified as segments from Arg-177 to Tyr-198 (2.6-kDa peptide) and from Ser-199 to Tyr-279 (9.1-kDa peptide). Their effects on actin polymerization induced by 50 or 100 mM KCl were studied by measuring the increase in viscosity by the falling ball method. The 2.6-kDa peptide decreased the rate of actin polymerization and increased the critical concentration for the polymerization. Based on the atomic model of the actin filament [Holmes, K.C., Popp, D., Gebhard, W., & Kabsch, W. (1990) Nature 347, 44-49], the peptide is presumed to bind to the barbed end of the actin filament and inhibit the polymerization. By assuming that the peptide affected the rate of association of the actin monomer to the end of the actin filament, well-fitting curves for the polymerization kinetics were calculated. Computer-assisted results indicated that the dissociation constant of the 2.6-kDa peptide for F-actin is 200 to 260 microM. In contrast, the 9.1-kDa peptide only slightly inhibited actin polymerization. These results suggest that the actin-actin interface in the region between Arg-177 and Tyr-198 has a stronger interaction than those between Ser-199 and Tyr-279. The amino acid sequence L-T-D-Y-L present in the 2.6-kDa segment is homologous to a common sequence in the F-actin capping domain of various actin-binding proteins.  相似文献   

17.
An improved DNase I inhibition assay for the filamentous actin (F-actin) and monomeric actin (G-actin) in brain cells has been developed. Unlike other methods, the cell lysis conditions and postlysis treatments, established by us, inhibited the temporal inactivation of actin in the cell lysate and maintained a stable F-actin/G-actin ratio for at least 4-5 h after lysis. The new procedure allowed separate quantitation of the noncytoskeletal F-actin in the Triton-soluble fraction (12,000 g, 10 min supernatant) that did not readily sediment with the Triton-insoluble cytoskeletal F-actin (12,000 g, 10 min pellet). We have applied this modified assay system to study the effect of hypothyroidism on different forms of actin using primary cultures of neurons derived from cerebra of neonatal normal and hypothyroid rats. Our results showed a 20% increase in the Triton-insoluble cytoskeletal F-actin in cultures from hypothyroid brain relative to normal controls. In the Triton-soluble fraction, containing the G-actin and the noncytoskeletal F-actin, cultures from hypothyroid brain showed a 15% increase in G-actin, whereas the F-actin remained unaltered. The 10% increase in total actin observed in this fraction from hypothyroid brain could be totally accounted for by the enhancement of G-actin. The mean F-actin/G-actin ratio in this fraction was about 30% higher in the cultures from normal brain compared to that of the hypothyroid system, which indicates that hypothyroidism tends to decrease the proportion of noncytoskeletal F-actin relative to G-actin.  相似文献   

18.
Cytochalasin D strongly inhibits the faster components in the reactions of actin filament depolymerization and elongation in the presence of 10 mM Tris-Cl-, pH 7.8, 0.2 mM dithiothreitol, 1 mM MgCl2, 0.1 mM CaCl2, and 0.2 mM ATP or ADP. Assuming an exclusive and total capping of the barbed end by the drug, the kinetic parameters derived at saturation by cytochalasin D refer to the pointed end and are 10-15-fold lower than at the barbed end. In ATP, the critical concentration increases with cytochalasin D up to 12-fold its value when both ends are free; as a result of the lowering of the free energy of nucleation by cytochalasin D, short oligomers of F-actin exist just above and below the critical concentration. Cytochalasin D interacts strongly with the barbed ends independently of the ADP-G-actin concentration (K = 0.5 nM-1). In contrast, the affinity of cytochalasin D decreases cooperatively with increasing ATP-G-actin concentration. These data are equally well accounted for by two different models: either cytochalasin D binds very poorly to ATP-capped filament ends whose proportion increases with actin concentration, or cytochalasin D binds equally well to ATP-ends and ADP-ends and also binds to actin dimers in ATP but not in ADP. A linear actin concentration dependence of the rate of growth was found at the pointed end, consistent with the virtual absence of an ATP cap at that end.  相似文献   

19.
The nucleotide state of actin (ATP, ADP-Pi, or ADP) is known to impact its interactions with other actin molecules upon polymerization as well as with multiple actin binding proteins both in the monomeric and filamentous states of actin. Recently, molecular dynamics simulations predicted that a sequence located at the interface of subdomains 1 and 3 (W-loop; residues 165–172) changes from an unstructured loop to a β-turn conformation upon ATP hydrolysis (Zheng, X., Diraviyam, K., and Sept, D. (2007) Biophys. J. 93, 1277–1283). This region participates directly in the binding to other subunits in F-actin as well as to cofilin, profilin, and WH2 domain proteins and, therefore, could contribute to the nucleotide sensitivity of these interactions. The present study demonstrates a reciprocal communication between the W-loop region and the nucleotide binding cleft on actin. Point mutagenesis of residues 167, 169, and 170 and their site-specific labeling significantly affect the nucleotide release from the cleft region, whereas the ATP/ADP switch alters the fluorescence of probes located in the W-loop. In the ADP-Pi state, the W-loop adopts a conformation similar to that in the ATP state but different from the ADP state. Binding of latrunculin A to the nucleotide cleft favors the ATP-like conformation of the W-loop, whereas ADP-ribosylation of Arg-177 forces the W-loop into a conformation distinct from those in the ADP and ATP-states. Overall, our experimental data suggest that the W-loop of actin is a nucleotide sensor, which may contribute to the nucleotide state-dependent changes in F-actin and nucleotide state-modulated interactions of both G- and F-actin with actin-binding proteins.  相似文献   

20.
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   

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