首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The formation of actin oligomers studied by analytical ultracentrifugation   总被引:2,自引:0,他引:2  
The small oligomers formed from Mg-G-actin under favorable conditions were studied by sedimentation velocity ultracentrifugation. The critical concentration of actin at pH 7.8 in the presence of 100 microM MgCl2 and 200 microM ATP was 12.5 +/- 2.8 microM. Under these conditions, about 15% of 7.5 microM Mg-actin was converted to oligomers of subunit size four to eight in 5 h at 20 degrees C. In 100 microM MgCl2 and no free ATP, the critical concentration was about 6.5 microM, and about 22% of 7.5 microM Mg-actin was converted to dimers in 80 min. There were no detectable higher oligomers or F-actin present in either case. As determined by the analysis of ATP hydrolysis, most, if not all, of the oligomer subunits contained ATP. When 28.5 microM actin was polymerized to steady state in 100 microM MgCl2 and 200 microM ATP, about 50% of the actin was present as F-actin, consistent with the critical concentration (approximately 12.5 microM), about 50% as oligomers as large as seven subunits, and only about 5% as monomers. When solutions containing oligomers were diluted the oligomers dissociated. Alternatively, when the MgCl2 concentration was raised to 1 mM, the solutions containing oligomers polymerized more rapidly than monomeric Mg-G-actin and to the same final steady state. These data are entirely consistent with the condensation-elongation model for helical polymerization proposed by Oosawa and Kasai (Oosawa, F., and Kasai, M. (1962) J. Mol. Biol. 4, 10-21) according to which, under certain conditions, substantial amounts of short linear and helical oligomers should be formed below the critical concentration and linear oligomers should coexist with monomers and F-actin at steady state.  相似文献   

2.
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   

3.
Mg2+, at submicromolar concentrations, is needed for the nucleation of actin [Maruyama (1981) J. Biol. Chem. 256, 1060-1062]. I show here that Mn2+ fulfils the same function. It binds to oligomers present in the ATP-G-actin solutions with a ratio of 2-3 Mn2+ ions per 100 actin monomers and with an association constant of 0.66 X 10(10) M-1 at pH 8.2 at 25 degrees C. The time course of the binding of Mn2+ to polymerizing actin is not affected by the initial concentration of the protein. Analysis of the distribution of the binding shows that, both in the large oligomeric species and in the polymers, 1 Mn2+ ion is bound for every 14-25 actin monomers, whereas in the smaller oligomeric species 1 Mn2+ ion is bound for every 4 actin monomers. The proposal is made that Mn2+ stabilizes actin nuclei and decreases the concentration of the monomers at the steady state. It is also proposed that, at least in some experimental conditions, the direct condensation of oligomers of intermediate length is an effective mechanism of F-actin formation.  相似文献   

4.
Modification of Lys-61 in actin with fluorescein-5-isothiocyanate (FITC) blocks actin polymerization [Burtnick, L. D. (1984) Biochim. Biophys. Acta 791, 57-62]. FITC-labelled actin recovered its ability to polymerize on addition of phalloidin. The polymers had the same characteristic helical thread-like structure as normal F-actin and the addition of myosin subfragment-1 to the polymers formed the characteristic arrowhead structure in electron microscopy. The polymers activated the ATPase activity of myosin subfragment-1 as efficiently as normal F-actin. These results indicate that Lys-61 is not directly involved in an actin-actin binding region nor in myosin binding site. From static fluorescence polarization measurements, the rotational relaxation time of FITC-labelled actin filaments was calculated to be 20 ns as the value reduced in water at 20 degrees C, while any rotational relaxation time of 1,5-IAEDANS bound to Cys-374 on F-actin in the presence of a twofold molar excess of phalloidin could not be detected by static polarization measurements under the same conditions. This indicates that the Lys-61 side chain is extremely mobile even in the filamentous structure. Fluorescence resonance energy transfer between the donor 1,5-IAEDANS bound to SH1 of myosin subfragment-1 and the acceptor fluorescein-5-isothiocyanate bound to Lys-61 of actin in the rigor complex was measured. The transfer efficiency was 0.39 +/- 0.05 which corresponds to the distance of 5.2 +/- 0.1 nm, assuming that the energy donor and acceptor rotate rapidly relative to the fluorescence lifetime and that the transfer occurs between a single donor and an acceptor.  相似文献   

5.
Phenylenebismaleimide has been used to form crosslinks between actin monomers [Knight, P. and Offer, G. (1978) Biochem. J. 175, 1023–1032]. We have purified a trimer of actin monomers as well as a dimer and a mixture of higher molecular weight oligomers. The trimer is much more effective than the dimer in enhancing the rate of polymerization while higher oligomers do not appear to be any more effective than the trimer. A lag in the polymerization process, as measured fluorescence enhancement of trace pyrene-actin, still occurs in the presence of trimers serving as the nuclei, suggesting that the mechanism for polymerization is more complex than nucleation followed by elongation.  相似文献   

6.
Formyl-met-leu-phe (fMLP) induces actin assembly in neutrophils; the resultant increase in F-actin content correlates with an increase in the rate of cellular locomotion at fMLP concentrations less than or equal to 10(-8) M (Howard, T.H., and W.H. Meyer, 1984, J. Cell Biol., 98:1265-1271). We studied the time course of change in F-actin content, F-actin distribution, and cell shape after fMLP stimulation. F-actin content was quantified by fluorescence activated cell sorter analysis of nitrobenzoxadiazole-phallacidin-stained cells (Howard, T.H., 1982, J. Cell Biol., 95(2, Pt. 2:327a). F-actin distribution and cell shape were determined by analysis of fluorescence photomicrographs of nitrobenzoxadiazole-phallacidin-stained cells. After fMLP stimulation at 25 degrees C, there is a rapid actin polymerization that is maximal (up to 2.0 times the control level) at 45 s; subsequently, the F-actin depolymerizes to an intermediate F-actin content 5-10 min after stimulation. The depolymerization of F-actin reflects a true decrease in F-actin content since the quantity of probe extractable from cells also decreases between 45 s and 10 min. The rate of actin polymerization (3.8 +/- 0.3-4.4 +/- 0.6% increase in F-actin/s) is the same for 10(-10) - 10(-6) M fMLP and the polymerization is inhibited by cytochalasin D. The initial rate of F-actin depolymerization (6.0 +/- 1.0-30 +/- 5% decrease in F-actin/min) is inversely proportional to fMLP dose. The F-actin content of stimulated cells at 45 s and 10 min is greater than control levels and varies directly with fMLP dose. F-actin distribution and cell shape also vary as a function of time after stimulation. 45 s after stimulation the cells are rounded and F-actin is diffusely distributed; 10 min after stimulation the cell is polarized and F-actin is focally distributed. These results indicate that actin polymerization and depolymerization follow fMLP stimulation in sequence, the rate of depolymerization and the maximum and steady state F-actin content but not the rate of polymerization are fMLP dose dependent, and concurrent with F-actin depolymerization, F-actin is redistributed and the cell changes shape.  相似文献   

7.
During inflammation, hydrogen peroxide, produced by polymorphonuclear leukocytes, provokes cell death mainly by disarranging filamentous (polymerized) actin (F-actin). To show the molecular mechanism(s) by which hydrogen peroxide could alter actin dynamics, we analyzed the ability of H2O2-treated actin samples to polymerize as well as the suitability of actin polymers (from oxidized monomers) to interact with cross-linking proteins. H2O2-treated monomeric (globular) actin (G-actin) shows an altered time course of polymerization. The increase in the lag phase and the lowering in both the polymerization rate and the polymerization extent have been evidenced. Furthermore, steady-state actin polymers, from oxidized monomers, are more fragmented than control polymers. This seems to be ascribable to the enhanced fragility of oxidized filaments rather than to the increase in the nucleation activity, which markedly falls. These facts; along with the unsuitability of actin polymers from oxidized monomers to interact with both filamin and alpha-actinin, suggest that hydrogen peroxide influences actin dynamics mainly by changing the F-actin structure. H2O2, via the oxidation of actin thiols (in particular, the sulfhydryl group of Cys-374), likely alters the actin C-terminus, influencing both subunit/subunit interactions and the spatial structure of the binding sites for cross-linking proteins in F-actin. We suggest that most of the effects of hydrogen peroxide on actin could be explained in the light of the "structural connectivity," demonstrated previously in actin.  相似文献   

8.
Length distribution of F-actin in Dictyostelium discoideum   总被引:9,自引:0,他引:9  
Inhibition of deoxyribonuclease I activity was used to assay the actin monomers and the pointed ends of actin filaments in lysates of Dictyostelium discoideum. The KD for the binding reaction was 0.2-0.3 nM. Total cellular actin was 93 microM in monomers (approximately 0.1 fmol/cell) of which roughly half was initially polymeric. Essentially all of the filamentous actin (F-actin) was readily pelleted in the microcentrifuge and was therefore presumed to be in the cytoskeleton. Free F-actin barbed ends, measured as pelletable [3H]cytochalasin B, numbered 1.8 x 10(5)/cell; nuclei for the polymerization of rabbit muscle globular (monomeric) actin numbered 2.0 x 10(5)/cell; and pointed ends, determined by their inhibition of deoxyribonuclease I, numbered 3.6 x 10(5)/cell. These values suggest that half the barbed ends might be occluded. On average, the filaments contained approximately 76 subunits and were therefore about 0.2 micron long. The distribution of their lengths was estimated from the time course of depolymerization following vast dilution. Three populations were defined. In one experiment, the smallest population contained 71% of the F-actin mass and 96% of the pointed ends; these filaments averaged 80 subunits or 0.22 microns in length. An intermediate population contained 14% of the F-actin mass and 3% of the filaments; these were roughly 460 subunits (1.3 microns) long. The largest population contained 15% of the F-actin mass in about 0.3% of the filaments; these were 13 microns in length, about the diameter of the cell. The numerous short filaments might populate a cortical mesh, while the long filaments might constitute endoplasmic bundles.  相似文献   

9.
Mitra K  Zhou D  Galán JE 《FEBS letters》2000,482(1-2):81-84
An essential step in the pathogenesis of Salmonella enterica infections is bacterial entry into non-phagocytic cells of the intestinal epithelium. Proteins injected by Salmonella into host cells stimulate cellular responses that lead to extensive actin cytoskeleton reorganization and subsequent bacterial uptake. One of these proteins, SipA, modulates actin dynamics by directly binding to F-actin. We have biophysically characterized a C-terminal fragment, SipA(446-684), which has previously been shown to retain activity. Our results show that SipA(446-684) exhibits an elongated shape with a predominantly helical conformation and predict the existence of a coiled-coil domain. We suggest that the protein is able to span two adjacent actin monomers in a filament and propose a model that is consistent with the observed effects of SipA(446-684) on actin dynamics and F-actin stability and morphology.  相似文献   

10.
When concentrated in mildly acidic solutions, bovine pancreatic ribonuclease (RNase A) forms long-lived oligomers including two types of dimer, two types of trimer, and higher oligomers. In previous crystallographic work, we found that the major dimeric component forms by a swapping of the C-terminal beta-strands between the monomers, and that the minor dimeric component forms by swapping the N-terminal alpha-helices of the monomers. On the basis of these structures, we proposed that a linear RNase A trimer can form from a central molecule that simultaneously swaps its N-terminal helix with a second RNase A molecule and its C-terminal strand with a third molecule. Studies by dissociation are consistent with this model for the major trimeric component: the major trimer dissociates into both the major and the minor dimers, as well as monomers. In contrast, the minor trimer component dissociates into the monomer and the major dimer. This suggests that the minor trimer is cyclic, formed from three monomers that swap their C-terminal beta-strands into identical molecules. These conclusions are supported by cross-linking of lysyl residues, showing that the major trimer swaps its N-terminal helix, and the minor trimer does not. We verified by X-ray crystallography the proposed cyclic structure for the minor trimer, with swapping of the C-terminal beta-strands. This study thus expands the variety of domain-swapped oligomers by revealing the first example of a protein that can form both a linear and a cyclic domain-swapped oligomer. These structures permit interpretation of the enzymatic activities of the RNase A oligomers on double-stranded RNA.  相似文献   

11.
The rate of filamentous actin (F-actin) depolymerization is proportional to the number of filaments depolarizing and changes in the rate are proportional to changes in filament number. To determine the number and length of actin filaments in polymorphonuclear leukocytes and the change in filament number and length that occurs during the increase in F-actin upon chemoattractant stimulation, the time course of cellular F-actin depolymerization in lysates of control and peptide-stimulated cells was examined. F-actin was quantified by the TRITC-labeled phalloidin staining of pelletable actin. Lysis in 1.2 M KCl and 10 microM DNase I minimized the effects of F-actin binding proteins and G-actin, respectively, on the kinetics of depolymerization. To determine filament number and length from a depolymerization time course, depolymerization kinetics must be limited by the actin monomer dissociation rate. Comparison of time courses of depolymerization in the presence (pointed ends free) or absence (barbed and pointed ends free) of cytochalasin suggested depolymerization occurred from both ends of the filament and that monomer dissociation was rate limiting. Control cells had 1.7 +/- 0.4 x 10(5) filaments with an average length of 0.29 +/- 0.09 microns. Chemo-attractant stimulation for 90 s at room temperature with 0.02 microM N-formylnorleucylleucylphenylalanine caused a twofold increase in F-actin and about a two-fold increase in the total number of actin filaments to 4.0 +/- 0.5 x 10(5) filaments with an average length of 0.27 +/- 0.07 microns. In both cases, most (approximately 80%) of the filaments were quite short (less than or equal to 0.18 micron). The length distributions of actin filaments in stimulated and control cells were similar.  相似文献   

12.
13.
α1-Proteinase inhibitor (antitrypsin) is a canonical example of the serpin family member that binds and inhibits serine proteases. The natural metastability of serpins is crucial to carry out structural rearrangements necessary for biological activity. However, the enhanced metastability of the mutant Z variant of antitrypsin, in addition to folding defect, may substantially contribute to its polymerization, a process leading to incurable serpinopathy. The metastability also impedes structural studies on the polymers. There are no crystal structures of Z monomer or any kind of polymers larger than engineered wild type (WT) trimer. Our understanding of polymerization mechanisms is based on biochemical data using in vitro generated WT oligomers and molecular simulations. Here we applied atomic force microscopy (AFM) to compare topography of monomers, in vitro formed WT oligomers, and Z type polymers isolated from transgenic mouse liver. We found the AFM images of monomers closely resembled an antitrypsin outer shell modeled after the crystal structure. We confirmed that the Z variant demonstrated higher spontaneous propensity to dimerize than WT monomers. We also detected an unexpectedly broad range of different types of polymers with periodicity and topography depending on the applied method of polymerization. Short linear oligomers of unit arrangement similar to the Z polymers were especially abundant in heat-treated WT preparations. Long linear polymers were a prominent and unique component of liver extracts. However, the liver preparations contained also multiple types of oligomers of topographies undistinguishable from those found in WT samples polymerized with heat, low pH or guanidine hydrochloride treatments. In conclusion, we established that AFM is an excellent technique to assess morphological diversity of antitrypsin polymers, which is important for etiology of serpinopathies. These data also support previous, but controversial models of in vivo polymerization showing a surprising diversity of polymer topography.  相似文献   

14.
During iron acquisition by the cell, complete homodimeric transferrin receptor 1 in an unknown state (R1) binds iron-loaded human serum apotransferrin in an unknown state (T) and allows its internalization in the cytoplasm. T also forms complexes with metals other than iron. Are these metals incorporated by the iron acquisition pathway and how can other proteins interact with R1? We report here a four-step mechanism for cobalt(III) transfer from CoNtaCO(3)(2-) to T and analyze the interaction of cobalt-loaded transferrin with R1. The first step in cobalt uptake by T is a fast transfer of Co(3+) and CO(3)(2-) from CoNtaCO(3)(2-) to the metal-binding site in the C-lobe of T: direct rate constant, k(1)=(1.1+/-0.1) x 10(6) M(-1) s(-1); reverse rate constant, k(-1)=(1.9+/-0.6) x 10(6) M(-1) s(-1); and equilibrium constant, K=1.7+/-0.7. This step is followed by a proton-assisted conformational change of the C-lobe: direct rate constant, k(2)=(3+/-0.3) x 10(6) M(-1) s(-1); reverse rate constant, k(-2)=(1.6+/-0.3) x 10(-2) s(-1); and equilibrium constant, K(2a)=5.3+/-1.5 nM. The two final steps are slow changes in the conformation of the protein (0.5 h and 72 h), which allow it to achieve its final thermodynamic state and also to acquire second cobalt. The cobalt-saturated transferrin in an unknown state (TCo(2)) interacts with R1 in two different steps. The first is an ultra-fast interaction of the C-lobe of TCo(2) with the helical domain of R1: direct rate constant, k(3)=(4.4+/-0.6)x10(10) M(-1) s(-1); reverse rate constant, k(-3)=(3.6+/-0.6) x 10(4) s(-1); and dissociation constant, K(1d)=0.82+/-0.25 muM. The second is a very slow interaction of the N-lobe of TCo(2) with the protease-like domain of R1. This increases the stability of the protein-protein adduct by 30-fold with an average overall dissociation constant K(d)=25+/-10 nM. The main trigger in the R1-mediated iron acquisition is the ultra-fast interaction of the metal-loaded C-lobe of T with R1. This step is much faster than endocytosis, which in turn is much faster than the interaction of the N-lobe of T with the protease-like domain. This can explain why other metal-loaded transferrins or a protein such as HFE-with a lower affinity for R1 than iron-saturated transferrin but with, however, similar or higher affinities for the helical domain than the C-lobe-competes with iron-saturated transferrin in an unknown state towards interaction with R1.  相似文献   

15.
Ca2+-calmodulin-dependent phosphorylation of myosin regulatory light chains by the catalytic COOH-terminal half of myosin light chain kinase (MLCK) activates myosin II in smooth and nonmuscle cells. In addition, MLCK binds to thin filaments in situ and F-actin in vitro via a specific repeat motif in its NH2 terminus at a stoichiometry of one MLCK per three actin monomers. We have investigated the structural basis of MLCK-actin interactions by negative staining and helical reconstruction. F-actin was decorated with a peptide containing the NH2-terminal 147 residues of MLCK (MLCK-147) that binds to F-actin with high affinity. MLCK-147 caused formation of F-actin rafts, and single filaments within rafts were used for structural analysis. Three-dimensional reconstructions showed MLCK density on the extreme periphery of subdomain-1 of each actin monomer forming a bridge to the periphery of subdomain-4 of the azimuthally adjacent actin. Fitting the reconstruction to the atomic model of F-actin revealed interaction of MLCK-147 close to the COOH terminus of the first actin and near residues 228-232 of the second. This unique location enables MLCK to bind to actin without interfering with the binding of any other key actin-binding proteins, including myosin, tropomyosin, caldesmon, and calponin.  相似文献   

16.
The distribution of actin in hamster sperm cells was studied during spermiogenesis, epididymal transit, in vitro capacitation and acrosome reaction by immunogold procedures using a polyclonal and two monoclonal antiactin antibodies. A predominant actin labeling (F-actin) was detected in the subacrosomal space of spermatids. Actin labeling was also observed under the plasma membrane of intercellular bridges and along the outer acrosomal membrane. In late spermatids there was both F-actin depolymerization and a loss of actin immunolabeling, thus suggesting a dispersion of G-actin monomers. No obvious labeling was evidenced in residual bodies. This pattern was observed with the three antiactin probes. In contrast, an actin labeling reappeared over the fibrous sheath of the flagellum in epididymal spermatozoa but only when the polyclonal antibody was used. Only one single actin reactive band was detected by immunoblotting of sperm extracts. Since the sperm tails were NBD phallacidin negative they were considered to contain either G-actin or actin oligomers rather than bundles of actin filaments. It is suggested that G-actin originating in the head of late spermatids was redistributed to the flagellum of epidymal spermatozoa. No further changes were noted after capacitation and acrosome reaction thus indicating no apparent effect on actin polymerization and distribution.  相似文献   

17.
The delta H associated with the thermal unfolding of G-actin has been determined by differential scanning calorimetry (DSC) to be 142 +/- 5 kcal/mol, with the Tm (melting temperature) at 57.2 +/- 0.5 degrees C, at pH 8.0 (heating rate 0.5 K/min). The transition is broad and cannot be treated as a single transition that mimics a two-state process, suggesting the existence of domains. Deconvolution is done to fit it into two quasi-independent two-state transitions. For F-actin, the transition is more cooperative, with a cooperative ratio (the ratio of van't Hoff enthalpy and calorimetric enthalpy) of 1.4, indicating intermonomer interaction. The delta H of the thermal unfolding of F-actin is 162 +/- 10 kcal/mol with a Tm at 67.0 +/- 0.5 degrees C. A state of G-actin similar to that of the heat-denatured form, designated D-actin, is obtained by removing tightly bound Ca2+ with EGTA. The DSC-detectable cooperative transition is completely lost when the free calcium concentration of the medium is 1 x 10(-11) M or lower, using a Ca2+/EGTA buffer system. However, circular dichroism (CD) shows that the helix content of actin, 32% in the G-form, is only partially reduced to 19% in this apo form. The CD spectrum and the helix content of the calcium-depleted actin are almost identical with those of the heat-denatured D form. This loss of 40% of the native helical content is irreversible in both cases. The remaining 60% of the native helical content cannot be further eliminated by heating to 95 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
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   

19.
A new model for the interaction of dystrophin with F-actin   总被引:10,自引:1,他引:9       下载免费PDF全文
The F-actin binding and cross-linking properties of skeletal muscle dystrophin-glycoprotein complex were examined using high and low speed cosedimentation assays, microcapillary falling ball viscometry, and electron microscopy. Dystrophin-glycoprotein complex binding to F-actin saturated near 0.042 +/- 0.005 mol/ mol, which corresponds to one dystrophin per 24 actin monomers. Dystrophin-glycoprotein complex bound to F-actin with an average apparent Kd for dystrophin of 0.5 microM. These results demonstrate that native, full-length dystrophin in the glycoprotein complex binds F-actin with some properties similar to those measured for several members of the actin cross-linking super- family of proteins. However, we failed to observe dystrophin- glycoprotein complex-induced cross-linking of F-actin by three different methods, each positively controlled with alpha-actinin. Furthermore, high speed cosedimentation analysis of dystrophin- glycoprotein complex digested with calpain revealed a novel F-actin binding site located near the middle of the dystrophin rod domain. Recombinant dystrophin fragments corresponding to the novel actin binding site and the first 246 amino acids of dystrophin both bound F- actin but with significantly lower affinity and higher capacity than was observed with purified dystrophin-glycoprotein complex. Finally, dystrophin-glycoprotein complex was observed to significantly slow the depolymerization of F-actin, Suggesting that dystrophin may lie along side an actin filament through interaction with multiple actin monomers. These data suggest that although dystrophin is most closely related to the actin cross-linking superfamily based on sequence homology, dystrophin binds F-actin in a manner more analogous to actin side-binding proteins.  相似文献   

20.
The three-dimensional structure of actin filaments decorated with the actin-binding domain of chick smooth muscle alpha-actinin (alpha A1-2) has been determined to 21-A resolution. The shape and location of alpha A1-2 was determined by subtracting maps of F-actin from the reconstruction of decorated filaments. alpha A1-2 resembles a bell that measures approximately 38 A at its base and extends 42 A from its base to its tip. In decorated filaments, the base of alpha A1-2 is centered about the outer face of subdomain 2 of actin and contacts subdomain 1 of two neighboring monomers along the long-pitch (two-start) helical strands. Using the atomic model of F-actin (Lorenz, M., D. Popp, and K. C. Holmes. 1993. J. Mol. Biol. 234:826-836.), we have been able to test directly the likelihood that specific actin residues, which have been previously identified by others, interact with alpha A1-2. Our results indicate that residues 86-117 and 350-375 comprise distinct binding sites for alpha-actinin on adjacent actin monomers.  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