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 375 毫秒
1.
Summary An actin-modulating protein has been isolated from the obliquely striated body wall muscle of the earthwormLumbricus terrestris. The isolation procedure included extraction in the absence of Ca2+ with subsequent ammonium sulfate fractionation followed by ion exchange chromatography and gel filtration. The purified modulator preparation appears on SDS-PAGE as a doublet of bands with molecular weights of 43 and 45 kDa. Both separately isolated components exhibit the same actin-modulating properties and therefore probably represent two isoforms. Substoichiometric amounts of modulator sever actin filaments and effectively promote nucleation of actin polymerization, which results in the formation of short actin filaments. Both effets are completely dependent on the presence of Ca2+ and activation of the modulator occurs in a narrow range of free Ca2+ concentrations around 10–6 mmol l–1. The modulator increases the critical concentration for actin polymerization, indicating that it binds to the fast polymerizing end of the actin filaments. The modulator forms a stoichiometric complex with two actin molecules whereby its ability to sever actin filaments is lost.The properties of the earthworm actin modulator are discussed in comparison with similar actinassociated proteins.Abbreviations BSA bovine serum albumin - DTE dithioerythritol - EWAM earthworm actin modulator - IEF isoelectric focusing - PAGE polyacrylamide gel electrophoresis - PMSF phenylmethylsulfonyl fluoride - SDS sodium dodecyl sulfate Dedicated to Professor Dr. K.-E. Wohlfarth-Bottermann, Bonn, on the occasion of his 65th birthday  相似文献   

2.
3.
Rabbit skeletal muscle actin has been labeled with the fluorescent sulfhydryl reagent iodoacetamidotetramethylrhodamine (rhodamine). The label is probably located on Cys-373 because prior treatment of the actin with N-ethylmaleimide prevents incorporation of rhodamine. When the rhodamine-actin is polymerized with MgSO4 or KCl, there is approximately a 1.5-fold increase in fluorescence. The change in fluorescence is correlated with incorporation of monomeric globular actin into polymer and can therefore be used as a quantitative measure of polymerization. Trace quantities of rhodamine-actin can be used to follow the kinetics of polymerization of unlabeled actin without disturbing the sample, and it is shown that the use of a capillary viscometer accelerates the rate of polymerization. Fluorescence photobleaching recovery experiments, which measure the diffusion of rhodamine-labeled actin, show that nondiffusible (apparent diffusion coefficient < 10?10 cm2/s) filaments appear during the polymerization process but that immediately after shearing these filaments are readily diffusible. These results demonstrate the destructive nature of hydrodynamic shear stress on polymerized actin.  相似文献   

4.
Summary We have observed the distribution of filamentous actin in growing hyphae of the oomyceteSaprolegnia ferax. The actin was stained by electroporating intact hyphae in the presence of 4×10–8 M rhodamine phalloidin. Hyphae quickly recovered from electroporation and showed an apical cap of densely packed actin filaments. The pores created by the electric shock resealed in 8–10min and within 1/2 h hyphae resumed growth and appeared normal. This technique allows us to observe actin arrays during growth and may prove to be a useful tool in determining the complex roles of actin in apical growth.Abbreviations RP rhodamine phalloidin - F-actin filamentous actin  相似文献   

5.
The effect of the type of metal ion (i.e., Ca2+, Mg2+, or none) bound to the high-affinity divalent cation binding site (HAS) of actin on filament assembly, structure, and dynamics was investigated in the absence and presence of the mushroom toxin phalloidin. In agreement with earlier reports, we found the polymerization reaction of G-actin into F-actin filaments to be tightly controlled by the type of divalent cation residing in its HAS. Moreover, novel polymerization data are presented indicating that LD, a dimer unproductive by itself, does incorporate into growing F-actin filaments. This observation suggests that during actin filament formation, in addition to the obligatory nucleation– condensation pathway involving UD, a productive filament dimer, a facultative, LD-based pathway is implicated whose abundance strongly depends on the exact polymerization conditions chosen. The “ragged” and “branched” filaments observed during the early stages of assembly represent a hallmark of LD incorporation and might be key to producing an actin meshwork capable of rapidly assembling and disassembling in highly motile cells. Hence, LD incorporation into growing actin filaments might provide an additional level of regulation of actin cytoskeleton dynamics. Regarding the structure and mechanical properties of the F-actin filament at steady state, no significant correlation with the divalent cation residing in its HAS was found. However, compared to native filaments, phalloidin-stabilized filaments were stiffer and yielded subtle but significant structural changes. Together, our data indicate that whereas the G-actin conformation is tightly controlled by the divalent cation in its HAS, the F-actin conformation appears more robust than this variation. Hence, we conclude that the structure and dynamics of the Mg–F-actin moiety within the thin filament are not significantly modulated by the cyclic Ca2+ release as it occurs in muscle contraction to regulate the actomyosin interaction via troponin.  相似文献   

6.
Actins are highly conserved proteins and key players in central processes in all eukaryotic cells. The two actins of the malaria parasite are among the most divergent eukaryotic actins and also differ from each other more than isoforms in any other species. Microfilaments have not been directly observed in Plasmodium and are presumed to be short and highly dynamic. We show that actin I cannot complement actin II in male gametogenesis, suggesting critical structural differences. Cryo-EM reveals that Plasmodium actin I has a unique filament structure, whereas actin II filaments resemble canonical F-actin. Both Plasmodium actins hydrolyze ATP more efficiently than α-actin, and unlike any other actin, both parasite actins rapidly form short oligomers induced by ADP. Crystal structures of both isoforms pinpoint several structural changes in the monomers causing the unique polymerization properties. Inserting the canonical D-loop to Plasmodium actin I leads to the formation of long filaments in vitro. In vivo, this chimera restores gametogenesis in parasites lacking actin II, suggesting that stable filaments are required for exflagellation. Together, these data underline the divergence of eukaryotic actins and demonstrate how structural differences in the monomers translate into filaments with different properties, implying that even eukaryotic actins have faced different evolutionary pressures and followed different paths for developing their polymerization properties.  相似文献   

7.
The polymerization of actin in solutions of purified calf spleen actin or profilactin (1–10 mg·ml-1) was followed by synchrotron radiation X-ray solution scattering. At the concentration used, polymerization of actin from profilactin or actin occurs without any lag phase. It is shown by a combination of solution scattering, model calculations and electron microscopy that contrary to the conclusions from previous viscometry studies, filaments form without any lag phase in profilactin solution but aggregate in bundles or networks. This phenomenon is independent of the method used to induce polymerization: slow temperature increase, temperature jump in the presence of polymerizing salts or fast mixing with salt. This aggregation explains the lower final viscosity levels, as compared to actin solutions, observed during the polymerization of actin from profilactin.  相似文献   

8.
9.
The binding of the enkephalin dimer [d-Ala2, Leu5-NH-CH2-]2 (DPE2) is characterized by (1) its high affinity for receptors on NG108-15 hybrid cells, the affinity constantK=4.7×109 M–1 is up to 8-fold that of monomers (0.6 to 1.0×109 M–1), and (2) a maximal binding capacity equal to one half that of the monomers. Kinetic studies showed that DPE2 binds with a 2-fold higher rate, k1=6.3×107 M–1min–1, than monomers (2.4 to 3.8×107 M–1min–1), and dissociates at a slower rate than monomers. Dissociation of DPE2 was consistently bi- or multiphasic but increased about 12% only after 3 hr of dissociation in the presence of a large excess of unlabeled enkephalin. The dissociation kinetics of monomers varied with enkephalin and experimental conditions used. Consistent with the value for the maximal binding capacity, the kinetic studies are interpreted in support of the hypothesis that DPE2 binds by cross-linking two subunits of one receptor.  相似文献   

10.
Polymerization of actin to form an elastic gel is one of the main mechanisms responsible for cellular motility. The particular problem addressed here stems from the need to model theoretically the growth of actin gel under controlled conditions, as observed in experiments. A biomimetic in vitro system which consists of a spherical latex bead, coated by the enzymatic protein ActA, and a reconstituted cytoplasm within which such beads are placed, induces polymerization of actin on the surface of the bead in the form of successive elastic thin spherical layers. Each newly formed layer pushes outward, and is pushed inward by, the already formed spherical layers which altogether constitute an elastic spherical shell of thickness h varying with time. Thus, a stress field is created in the shell which in turn affects the rate of polymerization as well as that of dissociation of actin gel. Given this bio-chemo-mechanical coupling, the accurate determination of the stress field becomes a subject of great importance for the understanding of the process, and it is the main objective of this work. The problem is addressed by first assuming appropriate constitutive laws for the actin gel elastic material, and then solving the only non-trivial stress equilibrium differential equation along the radial direction assuming spherical symmetry. A linear and a non-linear constitutive model for isotropic elasticity is used, appropriate for small and finite strains, respectively, and the solution is found in closed analytical forms in both cases. Two important conclusions are reached. First, the stress field depends strongly on the compressibility of the actin gel medium via the value of the Poisson ratio, for both linear and non-linear analysis. Second, the linear and non-linear solutions are very close for small strains, but they diverge progressively as the strains increase from small to large. Guided by available experimental data on the observed strain levels, the analytical results are illustrated by selected graphs of stress variation along the radial direction. At the end some comments and suggestions on the bio-chemo-mechanical coupling of actin gel growth and resorption are presented, where the role of properly defined joint isotropic invariants of stress and a unit vector along the predominant direction of free ends of actin filaments at the polymerization site is introduced.  相似文献   

11.
Summary Microelectrodes have been used for the measurement of oxygen tension in various biological systems.(Silver, 1987) Although not previously reported, microelectrodes allow the direct measurement of oxygen tension profiles within collagen gels containing entrapped hepatocytes (collagcn-hepatocyte gels). These oxygen tension profiles, along with hepatocyte oxygen consumption data, allowed the estimation of a diffusion coefficient for oxygen in collagen-hepatocyte gels, D g = 2.99 × 10–5 cm2/s.  相似文献   

12.
The mechanism of salt-induced actin polymerization involves the energetically unfavorable nucleation step, followed by filament elongation by the addition of monomers. The use of a bifunctional cross-linker, N,N′-(1,4-phenylene)dimaleimide, revealed rapid formation of the so-called lower dimers (LD) in which actin monomers are arranged in an antiparallel fashion. The filament elongation phase is characterized by a gradual LD decay and an increase in the yield of “upper dimers” (UD) characteristic of F-actin. Here we have used 90° light scattering, electron microscopy, and N,N′-(1,4-phenylene)dimaleimide cross-linking to reinvestigate relationships between changes in filament morphology, LD decay, and increase in the yield of UD during filament growth in a wide range of conditions influencing the rate of the nucleation reaction. The results show irregularity and instability of filaments at early stages of polymerization under all conditions used, and suggest that an earlier documented coassembling of LD with monomeric actin contributes to the initial disordering of the filaments rather than to the nucleation of polymerization. The effects of the type of G-actin-bound divalent cation (Ca2+/Mg2+), nucleotide (ATP/ADP), and polymerizing salt on the relation between changes in filament morphology and progress in G-actin-to-F-actin transformation show that ligand-dependent alterations in G-actin conformation determine not only the nucleation rate but also the kinetics of ordering of the filament structure in the elongation phase. The time courses of changes in the yield of UD suggest that filament maturation involves cooperative propagation of “proper” interprotomer contacts. Acceleration of this process by the initially bound MgATP supports the view that the filament-destabilizing conformational changes triggered by ATP hydrolysis and Pi liberation during polymerization are constrained by the intermolecular contacts established between MgATP monomers prior to ATP hydrolysis. An important role of contacts involving the DNase-I-binding loop and the C-terminus of actin is proposed.  相似文献   

13.
Type II par operons harness polymerization of the dynamically unstable actin-like protein ParM to segregate low-copy plasmids in rod-shaped bacteria. In this study, we use time-lapse fluorescence microscopy to follow plasmid dynamics and ParM assembly in Escherichia coli. Plasmids lacking a par operon undergo confined diffusion with a diffusion constant of 5 × 10−5 μm2/s and a confinement radius of 0.28 μm. Single par-containing plasmids also move diffusively but with a larger diffusion constant (4 × 10−4 μm2/s) and confinement radius (0.42 μm). ParM filaments are dynamically unstable in vivo and form spindles that link pairs of par-containing plasmids and drive them rapidly (3.1 μm/min) toward opposite poles of the cell. After reaching the poles, ParM filaments rapidly and completely depolymerize. After ParM disassembly, segregated plasmids resume diffusive motion, often encountering each other many times and undergoing multiple rounds of ParM-dependent segregation in a single cell cycle. We propose that in addition to driving segregation, the par operon enables plasmids to search space and find sister plasmids more effectively.  相似文献   

14.
1.To study proteins transported with actin in axons, we pulse-labeled motoneurons in the chicken sciatic nerve with [35S]methionine and, 1–20 days later, isolated actin and its binding proteins by affinity chromatography of Triton soluble nerve extracts on DNase I–Sepharose. The DNase I-purified proteins were electrophoresed on two-dimensional gels and the specific activity of the radioactively labeled protein spots was estimated by fluorography.2.In addition to actin, which binds specifically to DNase I, a small number of other proteins were labeled, including established actin monomer binding proteins and a protein of 36 kDa and pI 8.5. On the basis of its molecular mass, pI, amino acid composition, and immunostaining, the unrecognized protein was identified as the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH).3.The high-affinity binding of GAPDH to actin was confirmed by incubation of Triton-soluble nerve extracts with either mouse anti-GAPDH (or antiactin) and indirect immunomagnetic separation with Dynabeads covalently linked to sheep anti-mouse antibody. Analysis by one-dimensional gel electrophoresis and immunoblotting showed that actin and GAPDH were the main proteins isolated by these methods.4.Analysis of labeled nerves at 12 and 20 days after pulse labeling showed that GAPDH and actin were transported at the same rate, i.e., 3–5 mm/day, which corresponds to slow component b of axonal transport. These proteins were not associated with rapidly transported proteins that accumulated proximal to a ligation 7 cm from the spinal cord 9 hr after injection of radioactivity.5.Our results indicate that GAPDH and actin are transported as a complex in axons and raise the possibility that GAPDH could act as a chaperone for monomeric actin, translocating it to intraaxonal sites for exchange with or assembly into actin filaments. Alternatively, actin could be involved in translocating and anchoring GAPDH to specialized sites in axons and nerve terminals that require a source of ATP by glycolysis.  相似文献   

15.
This investigation explored the mechanism for inhibition of β2 integrin adhesion molecules when neutrophils are exposed to nitric oxide (NO). Roles for specific proteins were elucidated using chemical inhibitors, depletion with small inhibitory RNA, and cells from knock-out mice. Optimal inhibition occurs with exposures to a NO flux of ∼28 nmol/min for 2 min or more, which sets up an autocatalytic cascade triggered by activating type 2 nitric-oxide synthase (NOS-2) and NADPH oxidase (NOX). Integrin inhibition does not occur with neutrophils exposed to a NOX inhibitor (Nox2ds), a NOS-2 inhibitor (1400W), or with cells from mice lacking NOS-2 or the gp91phox component of NOX. Reactive species cause S-nitrosylation of cytosolic actin that enhances actin polymerization. Protein cross-linking and actin filament formation assays indicate that increased polymerization occurs because of associations involving vasodilator-stimulated phosphoprotein, focal adhesion kinase, and protein-disulfide isomerase in proximity to actin filaments. These effects were inhibited in cells exposed to ultraviolet light which photo-reverses S-nitrosylated cysteine residues and by co-incubations with cytochalasin D. The autocatalytic cycle can be arrested by protein kinase G activated with 8-bromo-cyclic GMP and by a high NO flux (∼112 nmol/min) that inactivates NOX.  相似文献   

16.
Tropomodulins (Tmods) are F-actin pointed end capping proteins that interact with tropomyosins (TMs) and cap TM-coated filaments with higher affinity than TM-free filaments. Here, we tested whether differences in recognition of TM or actin isoforms by Tmod1 and Tmod3 contribute to the distinct cellular functions of these Tmods. We found that Tmod3 bound ∼5-fold more weakly than Tmod1 to α/βTM, TM5b, and TM5NM1. However, surprisingly, Tmod3 was as effective as Tmod1 at capping pointed ends of skeletal muscle α-actin (αsk-actin) filaments coated with α/βTM, TM5b, or TM5NM1. Tmod3 only capped TM-coated αsk-actin filaments more weakly than Tmod1 in the presence of recombinant αTM2, which is unacetylated at its NH2 terminus, binds F-actin weakly, and has a disabled Tmod-binding site. Moreover, both Tmod1 and Tmod3 were similarly effective at capping pointed ends of platelet β/cytoplasmic γ (γcyto)-actin filaments coated with TM5NM1. In the absence of TMs, both Tmod1 and Tmod3 had similarly weak abilities to nucleate β/γcyto-actin filament assembly, but only Tmod3 could sequester cytoplasmic β- and γcyto-actin (but not αsk-actin) monomers and prevent polymerization under physiological conditions. Thus, differences in TM binding by Tmod1 and Tmod3 do not appear to regulate the abilities of these Tmods to cap TM-αsk-actin or TM-β/γcyto-actin pointed ends and, thus, are unlikely to determine selective co-assembly of Tmod, TM, and actin isoforms in different cell types and cytoskeletal structures. The ability of Tmod3 to sequester β- and γcyto-actin (but not αsk-actin) monomers in the absence of TMs suggests a novel function for Tmod3 in regulating actin remodeling or turnover in cells.  相似文献   

17.
Expression of proteins involved in the adhesion of Listeria monocytogenes to mammalian cells or in the intracellular life cycle of this bacterium, including listeriolysin O (LLO), ActA, Ami, and InlB, was used to compare two populations of L. monocytogenes strains. One of the populations comprised 300 clinical strains, and the other comprised 150 food strains. All strains expressed LLO, InlB, and ActA. No polymorphism was observed for LLO and InlB. Ami was detected in 283 of 300 human strains and in 149 of 150 food strains. The strains in which Ami was not detected were serovar 4b strains. Based on the molecular weights of the proteins detected, the strains were divided into two groups with Ami (groups Ami1 [75% of the strains] and Ami2 [21%]) and into four groups with ActA (groups ActA1 [52% of the strains], ActA2 [18%], ActA3 [30%], and ActA4 [one strain isolated from food]). Logistic regression showed that food strains were more likely to belong to group ActA3 than human strains (odds ratio [OR] = 2.90; P = 1 × 10−4). Of the strains isolated from patients with non-pregnancy-related cases of listeriosis, bacteremia was predominantly associated with group Ami1 strains (OR = 1.89; P = 1 × 10−2) and central nervous system infections were associated with group ActA2 strains (OR = 3.04; P = 1 × 10−3) and group ActA3 strains (OR = 3.91; P = 1 × 10−3).  相似文献   

18.
Effect of capping protein on the kinetics of actin polymerization   总被引:11,自引:0,他引:11  
Acanthamoeba capping protein increased the rate of actin polymerization from monomers with and without calcium. In the absence of calcium, capping protein also increased the critical concentration for polymerization. Various models were evaluated for their ability to predict the effect of capping protein on kinetic curves for actin polymerization under conditions where the critical concentration was not changed. Several models, which might explain the increased rate of polymerization from monomers, were tested. Two models which predicted the experimental data poorly were (1) capping protein was similar to an actin filament, bypassing nucleation, and (2) capping protein fragmented filaments. Three models in which capping protein accelerated, but did not bypass, nucleation predicted the data well. In the best one, capping protein resembled a nondissociable actin dimer. Several lines of evidence have supported the idea that capping protein blocks the barbed end of actin filaments, preventing the addition and loss of monomers [Cooper, J. A., Blum, J. D., & Pollard, T. D. (1984) J. Cell Biol. 99, 217-225; Isenberg, G. A., Aebi, U., & Pollard, T. D. (1980) Nature (London) 288, 455-459]. This mechanism was also supported here by the effect of capping protein on the kinetics of actin polymerization which was nucleated by preformed actin filaments. Low capping protein concentrations slowed nucleated polymerization, presumably because capping protein blocked elongation at barbed ends of filaments. High capping protein concentrations accelerated nucleated polymerization because of capping protein's ability to interact with monomers and accelerate nucleation.  相似文献   

19.
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   

20.
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   

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