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.
Dynamic light scattering was used to characterize the diffusion of monodisperse polystyrene latex spheres (PLS) of different sizes (55-, 105-, and 265-nm radii) in column-purified 0.65 mg/mL actin solutions polymerized with 100 mM KCl in the absence and presence of various amounts of the actin-binding protein gelsolin. The gelsolin and its interaction with actin was initially studied to ensure that the gelsolin could be used to produce filament populations with well-defined mean lengths. Measurements with PLS diffusion probes present showed, in the absence of gelsolin, that the effective local microviscosity in the actin solutions was 5-20 times that of water and that a large fraction of the PLS are trapped within the pores of the actin filament network, as found previously [J. Newman, K. L. Schick, & K. S. Zaner, (1989) Biopolymers 28, 1969-1980]. As the molar ratio of gelsolin to actin was increased, the diffusion coefficients of the PLS approached those in pure water while the degree of PLS trapping went to zero. Measurements of the dependence of the PLS diffusion coefficients on the ratio of actin concentration to the semidilute overlap concentration showed, for the smaller PLS, a transition occurring near the mean global overlap concentration. These results reflect the dissolution of the actin network as the gelsolin concentration is increased and illustrate the role of gelsolin/actin interactions in the control of macromolecular transport within the periphery of cells.  相似文献   

2.
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   

3.
We used actin filament bundles isolated from intestinal brush-border microvilli to nucleate the polymerization of pure muscle actin monomers into filaments. Growth rates were determined by electron microscopy by measuring the change in the length of the filaments as a function of time. The linear dependence of the growth rates on the actin monomer concentration provided the rate constants for monomer association and dissociation at the two ends of the growing filament. The rapidly growing ("barbed") end has higher association and dissociation rate constants than the slowly growing ("pointed") end. The values of these rate constants differ in 20 mM KCl compared with 75 mM KCl, 5 mM MgSO4. 2 microM cytochalasin B blocks growth entirely at the barbed end, apparently by reducing both association and dissociation rate constants to near zero, but inhibits growth at the pointed end to only a small extent.  相似文献   

4.
Increasing potassium chloride concentration from 0 to 100 mM and magnesium chloride from 0 to 2 mM show a parallel rate increase in polymerizing actin, whereas increasing calcium chloride concentration from 0 to 0.2 mM decreases the rate of polymerizing actin. The presence of alpha-actinin has little influence on the polymerization kinetics of actin under these conditions. Viscometric measurements indicate that the presence of various mono- and divalent cations, ionic strength, and alpha-actinin in combination are responsible for changes in the mechanical properties of solutions containing actin. The actin filament dynamic behavior is drastically reduced under these conditions as confirmed by quasi-elastic light scattering.  相似文献   

5.
Yu X  Carlsson AE 《Biophysical journal》2004,87(6):3679-3689
We study the kinetics of filament bundling by variable time-step Brownian-dynamics simulations employing a simplified attractive potential based on earlier atomic-level calculations for actin filaments. Our results show that collisions often cluster in time, due to memory in the random walk. The clustering increases the bundling opportunities. Small-angle collisions and collisions with short center-to-center distance are more likely to lead to bundling. Increasing the monomer-monomer attraction decreases the bundling time to a diffusional limit, which is determined by the capture cross-section and diffusion coefficients. The simulations clearly show that the bundling process consists of two sequential phases: rotation, by which two filaments align parallel to each other; and sliding, by which they maximize their contact length. Whether two filaments bundle or not is determined by the competition between rotation to a parallel state and escape. Increasing the rotational diffusion coefficient and attraction enhances rotation; decreasing attraction and increasing the translational diffusion coefficients enhance escape. Because of several competing effects, the filament length only affects the bundling time weakly.  相似文献   

6.
Actin filaments of different lengths were prepared by polymerizing actin in the presence of various concentrations of gelsolin, a protein which accelerates actin polymerization by stabilizing nuclei from which filaments grow and which binds to their fast growing ends. The lengths of the actin filaments following polymerization were measured by electron microscopy and showed that the number-average filament length agreed with the predicted length if each gelsolin molecule acted as a seed for the growth of an actin filament. The distribution of lengths was independent of the actin:gelsolin ratio and was similar to that of actin filaments polymerized in the absence of gelsolin (Lw/Ln = 1.8). The mobility of these filaments in solution was studied by quasielastic light scattering and by viscometry. The translational diffusion constant determined by quasielastic light scattering was in agreement with the infinite dilution values calculated from the dimensions and the distribution of lengths determined by electron microscopy for relatively short filament lengths. Under conditions where overlap of the rotational domains of the filaments would be expected to occur, the measured diffusion rates deviated from their predicted dilute solution values and the solution viscosity increased abruptly. The dependence of the diffusion constant and the solution viscosity on the length of the actin filaments can be explained in terms of a theory that describes the restraints on diffusion of independent rigid rods in semi-dilute solution. The results suggest that the rheology of actin filaments can be accounted for by steric restraints. The length of cytoplasmic actin filaments in some cell types is such that these steric constraints are significant and could produce large changes in physical properties with small changes in filament length.  相似文献   

7.
Yang L  Sept D  Carlsson AE 《Biophysical journal》2006,90(12):4295-4304
The formation of filopodia-like bundles from a dendritic actin network has been observed to occur in vitro as a result of branching induced by Arp2/3 complex. We study both the energetics and dynamics of actin filament bundling in such a network to evaluate their relative importance in bundle formation processes. Our model considers two semiflexible actin filaments fixed at one end and free at the other, described using a normal-mode approximation. This model is studied by both Brownian dynamics and free-energy minimization methods. Remarkably, even short filaments can bundle at separations comparable to their lengths. In the dynamic simulations, we evaluate the time required for the filaments to interact and bind, and examine the dependence of this bundling time on the filament length, the distance between the filament bases, and the cross-linking energy. In most cases, bundling occurs in a second or less. Beyond a certain critical distance, we find that the bundling time increases very rapidly with increasing interfilament separation and/or decreasing filament length. For most of the cases we have studied, the energetics results for this critical distance are similar to those obtained from dynamics simulations run for 10 s, suggesting that beyond this timescale, energetics, rather than kinetic constraints, determine whether or not bundling occurs. Over a broad range of conditions, we find that the times required for bundling from a network are compatible with experimental observations.  相似文献   

8.
Substoichiometric concentrations of cytochalasin D inhibited the rate of polymerization of actin in 0.5 mM MgCl2, increased its critical concentration and lowered its steady state viscosity. Stoichiometric concentrations of cytochalasin D in 0.5 mM MgCl2 and even substoichiometric concentrations of cytochalasin D in 30 mM KCl, however, accelerated the rate of actin polymerization, although still lowering the final steady state viscosity. Cytochalasin B, at all concentrations in 0.5 mM MgCl2 or in 30 mM KCl, accelerated the rate of polymerization and lowered the final steady state viscosity. In 0.5 mM MgCl2, cytochalasin D uncoupled the actin ATPase activity from actin polymerization, increasing the ATPase rate by at least 20 times while inhibiting polymerization. Cytochalasin B had a very much lower stimulating effect. Neither cytochalasin D nor B affected the actin ATPase activity in 30 mM KCl. The properties of cytochalasin E were intermediate between those of cytochalasin D and B. Cytochalasin D also stimulated the ATPase activity of monomeric actin in the absence of MgCl2 and KCl and, to a much greater extent, stimulated the ATPase activity of monomeric actin below its critical concentration in 0.5 mM MgCl2. Both above and below its critical concentration and in the presence and absence of cytochalasin D, the initial rate of actin ATPase activity, when little or no polymerization had occurred, was directly proportional to the actin concentration and, therefore, apparently was independent of actin-actin interactions. To rationalize all these data, a working model has been proposed in which the first step of actin polymerization is the conversion of monomeric actin-bound ATP, A . ATP, to monomeric actin-bound ADP and Pi, A* . ADP . Pi, which, like the preferred growing end of an actin filament, can bind cytochalasins.  相似文献   

9.
The diffusion coefficients of monodisperse polystyrene latex spheres in solutions of polymerized actin were measured using dynamic light scattering. Four different probes with radii R, ranging from 50 to 500 nm, were separately used in actin solutions with concentrations c, ranging from 1.5 to 21 microM, which had been polymerized with either 1 mM MgCl2, 1 mM CaCl2, or 100 mM KCl. Under all conditions, and at four different scattering angles in the range of 30 degrees-90 degrees, the measured average diffusion coefficients D of the probes were systematically smaller for samples of increased actin concentration or of increased probe radius. Control experiments indicated that the probes did not bind to the actin. These data for Mg2+- and Ca2+-polymerized actin agree and were found to be quite well summarized by the scaling relation D/D0 = exp[-alpha R delta c nu], where D0 is the measured diffusion coefficient of the probes in water (and, as also measured, in the starting actin solutions prior to polymerization with added salt), with values of delta = 0.73 +/- 0.05, nu = 1.08 +/- 0.09, and alpha = (1.1 +/- 0.6) x 10(-3) (with c in microM and R in nm). Data for KCl-polymerized actin show much more restricted diffusivities of the probes at comparable actin concentrations. Inhomogeneities in the solution are reflected in the "effective polydispersity" of the probe diffusion coefficients, which depend on local microviscosity differences.  相似文献   

10.
C M Cohen  S F Foley 《Biochemistry》1984,23(25):6091-6098
Ternary complex formation between the major human erythrocyte membrane skeletal proteins spectrin, protein 4.1, and actin was quantified by measuring cosedimentation of spectrin and band 4.1 with F-actin. Complex formation was dependent upon the concentration of spectrin and band 4.1, each of which promoted the binding of the other to F-actin. Simultaneous measurement of the concentrations of spectrin and band 4.1 in the sedimentable complex showed that a single molecule of band 4.1 was sufficient to promote the binding of a spectrin dimer to F-actin. However, the molar ratio of band 4.1/spectrin in the complex was not fixed, ranging from approximately 0.6 to 2.2 as the relative concentration of added spectrin to band 4.1 was decreased. A mole ratio of 0.6 band 4.1/spectrin suggests that a single molecule of band 4.1 can promote the binding of more than one spectrin dimer to an actin filament. Saturation binding studies showed that in the presence of band 4.1 every actin monomer in a filament could bind at least one molecule of spectrin, yielding ternary complexes with spectrin/actin mole ratios as high as 1.4. Electron microscopy of such complexes showed them to consist of actin filaments heavily decorated with spectrin dimers. Ternary complex formation was not affected by alteration in Mg2+ or Ca2+ concentration but was markedly inhibited by KCl above 100 mM and nearly abolished by 10 mM 2,3-diphosphoglycerate or 10 mM adenosine 5'-triphosphate. Our data are used to refine the molecular model of the red cell membrane skeleton.  相似文献   

11.
Actin polymerization was investigated using fluorescence probe N-(1-pyrenyl)iodoacetamide, which was bound covalently to reactive sulfhydryl group, Cys-373. Labeled actin in the bulk was 0.5 to 1% of total actin concentration. Actin polymerization at concentration 12 mM was started by addition of 20 mM KCl and 2 mM MgCl2. The label fluorescence was excited at 365 nm and registered at 386 nm. Under actin polymerization the label fluorescence increased almost 10 times. Two main phases may be distinguished in the process of actin polymerization: 1) monomer activation and nucleus (trimer) formation, 2) growth of actin filaments on the nuclei. In our experimental conditions, both for pure actin and for that with added annexin VI, the 1st phase continued for about 3 min and after that the 2nd phase was perfectly approximated by exponential dependence. An analysis of the exponential curves showed that actin monomer lifetime increased from 327 s, at annexin absence, to about 373 s at 0.7 microM annexin and more. Calculation of rate constants at two ends of growing actin filament suggests that annexin VI binds with pointed ("slow") end so that at sufficient annexin concentration the filament grows only on barbed ("fast") end. Our results, together with data of other researchers showing that annexin VI binds with the inner membrane surface of smooth muscle cell through Ca2+, may indicate that, at Ca2+ entering the cell, this annexin binds actin filament pointed ends to cell surface making it ready for the act of contraction.  相似文献   

12.
Fluorescence photobleaching recovery (FPR) was measured to determine the diffusion coefficient of fluorescein-labeled G-actin in low-salt buffer. The result obtained, 7.15 +/- 0.35 X 10(-7) cm2/s, is in good agreement with that computed from the molecular weight, partial specific volume, and sedimentation coefficient, but is higher than previously obtained values. It is demonstrated from theory that at low ionic strength, the electrostatic contribution to the intrinsic viscosity leads to an overestimate of the hydrodynamic eccentricity of G-actin. Data from FPR, sedimentation, and fluorescence polarization experiments all indicate that the true low-salt form of the actin monomer has an axial ratio less than or equal to 3.0. The G-F transformation of actin was also observed by measurement of FPR during the assembly phase, in the steady state, and in the presence of ligands such as cytochalasin and aldolase. Each FPR record in general yields three data: relative proportion of rapidly and slowly diffusing actin, diffusion coefficient for the high-mobility fraction, and a mean diffusion coefficient for the low-mobility fraction. A relation between the mean low-mobility diffusion coefficient and the number-average filament length is derived and applied to the analysis of FPR data. Under typical conditions, the average filament length was much greater than 10 micron in the steady state. Cytochalasin D was found to decrease filament length and total amount of filament proportionally; total filament number was not greatly affected. In all polymerizations of G-actin, the high-mobility material observed in situ was found to be essentially monomeric actin. Relatively stable oligomers of actin were separated by fractionating G-AF-actin by gel filtration in 50 microM MgCl2 at 4 degrees C. On the basis of the diffusion coefficient, we conclude that monomer and dimer constitute the major particle types present under these conditions. Sedimentation of labeled actin polymerized in 1.0 mM MgCl2 yielded a graded supernatant that contained actin oligomers significantly larger than the monomer.  相似文献   

13.
The actin-activated Mg2+-ATPase activity of smooth muscle myosin was measured in 85 mM KCl, 6 mM MgCl2 in the absence of tropomyosin. The activity was dependent on myosin concentration. Vmax increased as myosin concentration was increased, while the Ka (the apparent dissociation constant for actin) remained the same. The extent of filament formation was also correlated with myosin concentration and most of the myosin monomers existed in 10S conformation. These results suggest that myosin concentration influences the actin-activated Mg2+-ATPase activity by changing the 10S-6S-filaments equilibrium.  相似文献   

14.
A homodimer protein consisting of two 38,000 dalton peptides was isolated from a murine leukemia cell line (M1). The binding molar ratio of the 38K-dimer protein to purified skeletal muscle actin was saturated at 1:3, and when the 38K-dimer/actin ratio exceeded 1:12, gelation occurred. This gelation was completely inhibited by the presence of either 10 mM KCl or 20 mM NaCl. The protein induced actin filament bundling, which required a higher 38K-dimer/actin ratio and was not affected by the presence of monovalent cations. During the differentiation of Ml cells, the sensitivity of the 38K protein to monovalent cations was decreased; that is 20 mM KCl or 50 mM NaCl was required to inhibit the gelation by the 38K protein isolated from differentiated cells. On the other hand, the intracellular K+ content of Ml cells decreased from 70 +/- 5 mM to 18 +/- 3 mM, and Na+ increased from 10 +/- 5 mM to 40 +/- 10 mM during the differentiation. These findings suggest that the differentiation brought about conditions favourable for the 38K protein to induce actin gelation, and in turn, the locomotive and phagocytic activities which were induced only after differentiation in this cell line.  相似文献   

15.
Assembly of Acanthamoeba actin, of which trace quantities had been labeled with 5-(iodoacetamido)-fluorescein, was quantified using the modulation detection method of fluorescence photobleaching recovery (FPR). This technique permits explicit determination of the fraction of labeled actin incorporated into filaments and the translational diffusion coefficients of the filaments, from which filament length can be calculated. Addition of Acanthamoeba profilin in molar ratios to actin of about 1.1:1 and 2.3:1 retarded the initial kinetics of assembly (induced by addition of 2mM Mg+2) and reduced the fraction of actin incorporated into filaments. The diffusion coefficients of filaments formed were greatly changed by the presence of profilin at short times, but the differences became increasingly smaller at longer times. After 26 hr. the filaments formed in 1.1:1 profilin were about 12% shorter and in 2.3:1 profilin were about 20% shorter than filaments formed by actin alone under the same conditions.  相似文献   

16.
Hu X  Kuhn JR 《PloS one》2012,7(2):e31385
We reconstructed cellular motility in vitro from individual proteins to investigate how actin filaments are organized at the leading edge. Using total internal reflection fluorescence microscopy of actin filaments, we tested how profilin, Arp2/3, and capping protein (CP) function together to propel thin glass nanofibers or beads coated with N-WASP WCA domains. Thin nanofibers produced wide comet tails that showed more structural variation in actin filament organization than did bead substrates. During sustained motility, physiological concentrations of Mg(2+) generated actin filament bundles that processively attached to the nanofiber. Reduction of total Mg(2+) abolished particle motility and actin attachment to the particle surface without affecting actin polymerization, Arp2/3 nucleation, or filament capping. Analysis of similar motility of microspheres showed that loss of filament bundling did not affect actin shell formation or symmetry breaking but eliminated sustained attachments between the comet tail and the particle surface. Addition of Mg(2+), Lys-Lys(2+), or fascin restored both comet tail attachment and sustained particle motility in low Mg(2+) buffers. TIRF microscopic analysis of filaments captured by WCA-coated beads in the absence of Arp2/3, profilin, and CP showed that filament bundling by polycation or fascin addition increased barbed end capture by WCA domains. We propose a model in which CP directs barbed ends toward the leading edge and polycation-induced filament bundling sustains processive barbed end attachment to the leading edge.  相似文献   

17.
Podlubnaia ZA  Nowak E 《Biofizika》2006,51(5):804-809
It was shown by electron microscopy that PEP33 a synthetic C-terminal peptide of the thymus hormone thymopoietin, formed bundles of actin 'filaments in the presence of 0.1 M KCl. The structure of PEP33 aggregates localizated in the bundles between actin filaments is very similar to that of aggregates observed in samples of pure PEP33. No changes were revealed in the structure of G-actin in the presence of PEP33. A similar, but a weaker bundling effect of thymopentin (PEP5) was also found. It forms bundles of actin filaments of small size. Further studies can shed light on the physiological importance of actin filament aggregation with the peptides of thymopoietin, the systematic release of which from the thymus produces the phenomena characteristic for the serious neuromuscular disease myasthenia gravis.  相似文献   

18.
A Suzuki  M Yamazaki  T Ito 《Biochemistry》1989,28(15):6513-6518
A high molecular weight inert molecule, poly(ethylene glycol) (PEG), or a soluble protein, ovalbumin, causes parallel bundles of actin filaments in a crystalline-like structure under physiological conditions of ionic compositions and pH. The bundle formation depends on the molecular weight of PEG, and a larger molecular weight of PEG can make the bundle at a lower concentration. Actin bundle formation has a discrete dependence on the concentration of PEG. The light scattering following PEG-induced bundle formation increased abruptly at 4.5% (w/w) PEG 6000, while at concentrations less than or equal to 4.0% (w/w) no increase was observed. Labeling actin filaments with heavy meromyosin indicated that the polarity of the filament in the bundle is random. The PEG-induced bundle formation depends on the ionic strength of the solutions and also the concentration of the filament, showing that a higher concentration of PEG was required at lower ionic strength or a lower concentration of the filament. The results described above cannot be explained on the basis of the postulation that the direct binding of PEG molecules to the actin filaments may cause bundle formation. Alternatively, the mechanism can be explained reasonably by the theory of osmoelastic coupling based on preferential exclusion of PEG molecules from the filament surface. High molecular weight molecules such as PEG should be preferentially excluded from the region adjacent to the actin filaments (exclusion layer) by steric hindrance, thereby making imbalance of osmolarity between the bulk and the exclusion layer. This imbalance puts an osmotic stress on the actin filament.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Yu X  Carlsson AE 《Biophysical journal》2003,85(6):3532-3543
An Ising-like counterion-binding model is developed and solved by a mean-field method. For G-actin, the calculated affinity constants of all the binding sites ranging from loose to tight binding match the experimental data. The model is used to calculate the interaction energy between two F-actin filaments. Within a certain counterion concentration range, a rapidly decaying attractive force between two parallel filaments is produced not only by the correlation of the counterion distributions on the two filaments, but also by the correlation of the configurations of the two filaments with fixed counterion positions, which has been ignored in previous calculations. The bundling energy depends strongly on the configuration of the filaments. Upon bundling, the tightly bound counterion site is not affected, but the medium and loosely bound ones are. The model reproduces the observed minimal divalent counterion concentration for bundling, and naturally predicts the resolubilization of bundles which is seen in recent experiments. At the optimal counterion concentration, we obtain a bundling energy of approximately -0.01 eV per monomer along the filament. The counterion valence strongly affects the optimal counterion concentration, but has only minor effects on the optimal bundling energy. We show that the attractive potential between filaments can be simplified as the sum of interactions between their monomers. This simplification makes it possible to calculate the exact free energy of a two-F-actin-filament system. We are thus able to probe the effects of filament length on F-actin bundling and obtain a critical length for bundling of 59 monomers at 1 microM monomer concentration and pH=7.2.  相似文献   

20.
M Oosawa  K Maruyama 《FEBS letters》1987,213(2):433-437
Phalloidin (2 mol per mol actin)-treated pyrenyl F-actin showed a critical concentration of 1.8 microM in the presence of 10 mM KCl, 0.2 mM ADP, and 5 mM Tris-HCl buffer, pH 8.0 at 25 degrees C. The filament weight concentration did not change at all during and after sonication, yet degrees of flow birefringence increased and the filament number concentration decreased after the termination of sonication. The latter changes were not affected by EDTA, but inhibited by beta-actinin. These observations suggest that reannealing of short pieces of phalloidin-treated actin filaments fragmented during sonication takes place during recovery after sonication.  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