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.
F-actin at steady state in the presence of ATP partially depolymerized to a new steady state upon mechanical fragmentation. The increase in critical concentration with the number concentration of filaments has been quantitatively studied. The data can be explained by a model in which the preferred pathway for actin association-dissociation reactions at steady state in the presence of ATP involves binding of G-actin . ATP to filaments, ATP hydrolysis, and dissociation of G-actin . ADP which is then slowly converted to G-actin . ATP. As a consequence of the slow exchange of nucleotide on G-actin, the respective amounts of G-actin . ATP and G-actin . ADP coexisting with F-actin at steady state depend on the filament number concentration. G-actin coexisting with F-actin at zero number concentration of filaments would then consist of G-actin . ATP only, while the critical concentration obtained at infinite number of filaments would be that for G-actin . ADP. Values of 0.35 and 8 microM, respectively, were found for these two extreme critical concentrations for skeletal muscle actin at 20 degrees C, pH 7.8, 0.1 mM CaCl2, 1 mM MgCl2, and 0.2 mM ATP. The same value of 8 microM was directly measured for the critical concentration of G-actin . ADP polymerized in the presence of ADP and absence of ATP, and it was unaffected by fragmentation. These results have important implications for experiments in which critical concentrations are compared under conditions that change the filament number concentrations.  相似文献   

2.
T Chen  M Haigentz  E Reisler 《Biochemistry》1992,31(11):2941-2946
The effect of myosin on the structure of two sequences on G-actin, a loop between residues 39 and 52 and a segment between residues 61 and 69 from the NH2-terminus, was probed by limited proteolytic digestions of G-actin in the presence of the myosin subfragment 1 isozyme S-1(A2). Under the experimental conditions of this work, no polymerization of actin was induced by S-1(A2) [Chen & Reisler (1991) Biochemistry 30, 4546-4552]. S-1(A2) did not change the rates of subtilisin and chymotryptic digestion of G-actin at loop 39-52. In contrast to this, the second protease-sensitive region on G-actin, segment 61-69, was protected strongly by S-1(A2) from tryptic cleavage. The minor if any involvement of loop 39-52 in S-1 binding was confirmed by determining the binding constants of S-1(A2) for pyrene-labeled G-actin (1.2 x 10(6) M-1), subtilisin-cleaved pyrenyl G-actin (0.3 x 10(6) M-1), and DNase I-pyrenyl G-actin complexes (0.3 x 10(6) M-1). Consistent with this, the activity of DNase I, which binds to actin loop 39-52 [Kabsch et al. (1990) Nature 347, 37-44], was inhibited almost equally well by actin in the presence and absence of S-1(A2). These results confirm the observation that DNase I and S-1(A2) bind to distinct sites on actin [Bettache et al. (1990) Biochemistry 29, 9085-9091] and demonstrate myosin-induced changes in segment 61-69 of G-actin.  相似文献   

3.
The interaction between G-actin and myosin subfragment-1 (S1) has been monitored by pyrenyl-actin fluorescence and light scattering. In low ionic strength buffer and in the absence of ATP the polymerization of G-actin induced by myosin subfragment-1 is preceded by the formation of binary GS and ternary G2S complexes in which S1 interacts tightly in rapid equilibrium (K greater than 10(7) M-1) with one and two G-actin molecules, respectively. Pyrenyl fluorescence of G-actin is enhanced 4-fold in GS and 3-fold in G2S. At concentrations of G-actin and S1 in the micromolar range and above, G2S is the predominant species at G-actin/S1 ratios equal to or greater than 1. The isomer of myosin subfragment-1 carrying the A1 light chain, S1(A1), forms a tighter ternary complex than the isomer S1(A2). Actin-bound ATP is not hydrolyzed upon formation of GS and G2S. In the presence of one molar equivalent or more of myosin subfragment-1/mol of G-actin, in low ionic strength buffer containing no nucleotides, G-actin polymerizes faster in the presence of S1(A1) than in the presence of S1(A2). The interaction of S1 with G-actin is inhibited by the binding of ATP or ADP to S1, ATP having a higher affinity for S1 than ADP. The possible structural similarity of the G2S complex to the F-acto-S1 complex in the rigor state and the potential significance of a ternary (actin)2-myosin interaction for actomyosin-based motility are discussed.  相似文献   

4.
V. N. Umetskaya 《Biophysics》2016,61(4):585-590
NMR proton spectra were recorded in the range of proton resonance in the nucleotide aromatic ring of monomeric ATP–G-actin and the Mg2+–ATP–G-actin solutions in D2O to study the mechanism of ATP–G-actin hydrolysis and its role in F-actin formation in Mg2+-containing solutions. The experimental data show variations in the proton chemical shifts of the H2 and H8 peaks and splitting of the H8 resonance peak of G-actin-bound ATP adenine caused by interaction with magnesium dication. The observed variations in spectra are explained by hydrolysis of monomeric ATP–G-actin to ADP–G-actin, which is regarded as the initial stage of the G-actin to F-actin transformation.  相似文献   

5.
Toxoplasma gondii ADF (TgADF) belongs to a functional subtype characterized by strong G-actin sequestering activity and low F-actin severing activity. Among the characterized ADF/cofilin proteins, TgADF has the shortest length and is missing a C-terminal helix implicated in F-actin binding. In order to understand its characteristic properties, we have determined the solution structure of TgADF and studied its backbone dynamics from 15N-relaxation measurements. TgADF has conserved ADF/cofilin fold consisting of a central mixed β-sheet comprised of six β-strands that are partially surrounded by three α-helices and a C-terminal helical turn. The high G-actin sequestering activity of TgADF relies on highly structurally and dynamically optimized interactions between G-actin and G-actin binding surface of TgADF. The equilibrium dissociation constant for TgADF and rabbit muscle G-actin was 23.81 nM, as measured by ITC, which reflects very strong affinity of TgADF and G-actin interactions. The F-actin binding site of TgADF is partially formed, with a shortened F-loop that does not project out of the ellipsoid structure and a C-terminal helical turn in place of the C-terminal helix α4. Yet, it is more rigid than the F-actin binding site of Leishmania donovani cofilin. Experimental observations and structural features do not support the interaction of PIP2 with TgADF, and PIP2 does not affect the interaction of TgADF with G-actin. Overall, this study suggests that conformational flexibility of G-actin binding sites enhances the affinity of TgADF for G-actin, while conformational rigidity of F-actin binding sites of conventional ADF/cofilins is necessary for stable binding to F-actin.  相似文献   

6.
Trypsin and chymotrypsin were used as probes of conformation of G-actin molecule. The pattern of fragments produced has been analyzed by sodium dodecyl sulfate gel electrophoresis. G-actin is known to be nonrefractory to proteolysis [Jacobson, G.R., and Rosenbusch, J.P. (1976) Proc. Natl. Acad. Sci. U.S.A. 73, 2742-2746]. It is really true that G-actin is cut easily into a 33-kDa fragment by trypsin or chymotrypsin, but only when free ATP is present in the medium. After the removal of free ATP from the medium, G-actin became more refractory to proteolysis. The amounts of degradation of G-actin depended on the ATP concentration in the medium with saturating at about 0.5 mM. epsilon-ADP also had the effect and its fluorescence spectrum was changed on the addition of G-actin. After the removal of free ATP, G-actin still bound 1 mol/mol of ATP. So, the present results suggest the presence of a second ATP interaction site on G-actin and that ATP interaction at this site induces conformational changes in G-actin molecule.  相似文献   

7.
The first step in the polymerisation of actin   总被引:7,自引:0,他引:7  
In the presence of certain cations (e.g. K+ or Mg2+) actin polymerizes. Below a certain concentration (the critical concentration) the monomer G-actin does not polymerize on the addition of K+ or Mg2+. However, the proteolysis experiments of Rich and Estes [J. Mol. Biol. 104, 777--792 (1976)] strongly suggest that cations induce a change in conformation of G-actin leading to a novel form of actin, G*-actin. This conformational change may be the first step in the polymerization of actin. We have studied G*-actin induced by K+, by difference spectroscopy. We show that G*-actin is a monomer and we confirm that the bound ATP is not cleaved. We also studied the G-actin in equilibrium with G*-actin equilibrium at 4 degrees C as a function of K+ or Mg2+ concentration. With KCl, the transformation can be accounted for as a screening effect. The effect of Mg2+ is more specific and the change in conformation of the G-actin could result from the binding of two or three Mg2+ ions/molecule. We suggest that the G-actin in equilibrium with G*-actin transformation results from the neutralization of a polyanionic region on the actin surface and that this region could be the highly negatively charged N terminus.  相似文献   

8.
Smooth muscle G-actin from chicken gizzard and Physarum plasmodium G-actin both interact with DNase I and form 1 : 1 complexes. These complexes were crystallized by using polyethylene glycol 6000 as a precipitant. Both crystals belong to the same orthorhombic space group P2(1)2(1)2(1). The cell dimensions of chicken gizzard G-actin.DNase I complex are a=42.00 +/- 0.07 A, b=225.3 +/- 0.4 A, and c=77.4 +/- 0.1 A, while those of Physarum G-actin.DNase I complex are a=42 A, b=221 A, and c=77 A.  相似文献   

9.
To understand the intracellular role of G-actin concentration in stimulus-induced actin assembly and lamellipodium extension during cell migration, we developed a novel technique for quantifying spatiotemporal changes in G-actin concentration in live cells, consisting of sequential measurements of fluorescent decay after photoactivation (FDAP) of Dronpa-labeled actin. Cytoplasmic G-actin concentrations decreased by ~40% immediately after cell stimulation and thereafter the cell area extended. The extent of stimulus-induced G-actin loss and cell extension correlated linearly with G-actin concentration in unstimulated cells, even at concentrations much higher than the critical concentration of actin filaments, indicating that cytoplasmic G-actin concentration is a critical parameter for determining the extent of stimulus-induced G-actin assembly and cell extension. Multipoint FDAP analysis revealed that G-actin concentration in lamellipodia was comparable to that in the cell body. We also assessed the cellular concentrations of free G-actin, profilin- and thymosin-β4-bound G-actin, and free barbed and pointed ends of actin filaments by model fitting of jasplakinolide-induced temporal changes in G-actin concentration.  相似文献   

10.
Snake venom cardiotoxin showed the ability to induce polymerization of G-actin from rabbit skeletal muscle in a low ionic strength buffer composed of 0.2 mM CaCl2/0.2 mM ATP/0.5 mM mercaptoethanol/2.0 mM Tris-HCl, pH 8.0. The activity was enhanced greatly when 0.4 mM MgCl2 was present in the buffer and could be inhibited if G-actin was preincubated with deoxyribonuclease I. Furthermore, the DNAase could also partially depolymerize actin polymer previously formed by the interaction of G-actin with the toxin.  相似文献   

11.
F-actin, a helical polymer formed by polymerization of the monomers (G-actin), plays crucial roles in various aspects of cell motility. Flexibility of F-actin has been suggested to be important for such a variety of functions. Understanding the flexibility of F-actin requires characterization of a hierarchy of dynamical properties, from internal dynamics of the actin monomers through domain motions within the monomers and relative motions between the monomers within F-actin to large-scale motions of F-actin as a whole. As a first step toward this ultimate purpose, we carried out elastic incoherent neutron scattering experiments on powders of F-actin and G-actin hydrated with D2O and characterized the internal dynamics of F-actin and G-actin. Well established techniques and analysis enabled the extraction of mean-square displacements and their temperature dependence in F-actin and in G-actin. An effective force constant analysis with a model consisting of three energy states showed that two dynamical transitions occur at ∼150 K and ∼245 K, the former of which corresponds to the onset of anharmonic motions and the latter of which couples with the transition of hydration water. It is shown that behavior of the mean-square displacements is different between G-actin and F-actin, such that G-actin is “softer” than F-actin. The differences in the internal dynamics are detected for the first time between the different structural states (the monomeric state and the polymerized state). The different behavior observed is ascribed to the differences in dynamical heterogeneity between F-actin and G-actin. Based on structural data, the assignment of the differences observed in the two samples to dynamics of specific loop regions involved in the polymerization of G-actin into F-actin is proposed.  相似文献   

12.
Intramonomer fluorescence energy transfer between the donor epsilon-ATP bound to the nucleotide-binding site and the acceptor 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole bound to Cys-373 in G-actin was measured by steady-state fluorimetry. Assuming for the orientation factor its dynamic limit K2 = 2/3, the donor and acceptor distance in a G-actin molecule was calculated to be about 3 nm. The intermonomer energy transfer in F-actin occurring between the donor bound to an actin monomer and the acceptor bound to the nearest-neighbour actin monomer was also measured and the distance was calculated to be about 4 nm. The kinetics of the actin polymerization process was studied by following the decrease in fluorescence intensity upon addition of salts to G-actin solution. The initial velocity of the fluorescence intensity change was proportional to the square of the initial G-actin concentration. The temperature dependence of the velocity was proportional to the square of the initial G-actin concentration. The temperature dependence of the velocity was proportional to exp(-10/RT). These results indicated that the initial fluorescence intensity change corresponds to monomer-dimer transformation and its activation enthalpy was 10 kcal/mol.  相似文献   

13.
Decameric vanadate species (V10) inhibit the rate and the extent of G-actin polymerization with an IC50 of 68+/-22 microM and 17+/-2 microM, respectively, whilst they induce F-actin depolymerization at a lower extent. On contrary, no effect on actin polymerization and depolymerization was detected for 2mM concentration of "metavanadate" solution that contains ortho and metavanadate species, as observed by combining kinetic with (51)V NMR spectroscopy studies. Although at 25 degrees C, decameric vanadate (10 microM) is unstable in the assay medium, and decomposes following a first-order kinetic, in the presence of G-actin (up to 8 microM), the half-life increases 5-fold (from 5 to 27 h). However, the addition of ATP (0.2mM) in the medium not only prevents the inhibition of G-actin polymerization by V10 but it also decreases the half-life of decomposition of decameric vanadate species from 27 to 10h. Decameric vanadate is also stabilized by the sarcoplasmic reticulum vesicles, which raise the half-life time from 5 to 18h whereas no effects were observed in the presence of phosphatidylcholine liposomes, myosin or G-actin alone. It is proposed that the "decavanadate" interaction with G-actin, favored by the G-actin polymerization, stabilizes decameric vanadate species and induces inhibition of G-actin polymerization. Decameric vanadate stabilization by cytoskeletal and transmembrane proteins can account, at least in part, for decavanadate toxicity reported in the evaluation of vanadium (V) effects in biological systems.  相似文献   

14.
N Bettache  R Bertrand  R Kassab 《Biochemistry》1990,29(38):9085-9091
We have investigated various structural and interaction properties of maleimidobenzoyl-G-actin (MBS-actin), a new, internally cross-linked G-actin derivative that does not exhibit, at moderate protein concentration, the salt--and myosin subfragment 1 (S-1)-induced polymerizations of G-actin and reacts reversibly and covalently in solution with S-1 at or near the F-actin binding region of the heavy chain (Bettache, N., Bertrand, R., & Kassab, R. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 6028-6032). The far-ultraviolet CD spectrum and alpha-helix content of the MBS-actin were identical with those displayed by native G-actin. 45Ca2+ measurements showed the same content of tightly bound Ca2+ in MBS-actin as in G-actin and the EDTA treatment of the modified protein promoted the same red shift of the intrinsic fluorescence spectrum as observed with native G-actin. Incubation of concentrated MBS-actin solutions with 100 mM KCl + 5 mM MgCl2 led to the polymerization of the actin derivative when the critical monomer concentration reached 1.6 mg/mL, at 25 degrees C, pH 8.0. The MBS-F-actin formed activated the Mg2(+)-ATPase of S-1 to the same extent as native F-actin. The MBS-G-actin exhibited a DNase I inhibitor activity very close to that found with native G-actin and was not to be at all affected by its specific covalent conjugation to S-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
β-Thymosin (βT) and WH2 domains are widespread, intrinsically disordered actin-binding peptides that display significant sequence variability and different regulations of actin self-assembly in motile and morphogenetic processes. Here, we reveal the structural mechanisms by which, in their 1:1 stoichiometric complexes with actin, they either inhibit assembly by sequestering actin monomers like Thymosin-β4, or enhance motility by directing polarized filament assembly like Ciboulot βT. We combined mutational, functional or structural analysis by X-ray crystallography, SAXS (small angle X-ray scattering) and NMR on Thymosin-β4, Ciboulot, TetraThymosinβ and the long WH2 domain of WASP-interacting protein. The latter sequesters G-actin with the same molecular mechanisms as Thymosin-β4. Functionally different βT/WH2 domains differ by distinct dynamics of their C-terminal half interactions with G-actin pointed face. These C-terminal interaction dynamics are controlled by the strength of electrostatic interactions with G-actin. At physiological ionic strength, a single salt bridge with actin located next to their central LKKT/V motif induces G-actin sequestration in both isolated long βT and WH2 domains. The results open perspectives for elucidating the functions of βT/WH2 domains in other modular proteins.  相似文献   

16.
Polymerization of G-actin to F-actin was indicated by an increase in light-scattering intensity after the addition of a heptapeptide (Ile-Arg-Ile-Cys(MT)-Arg-Lys-Gly-OEt), an analog of the actin-binding S-site on S-1 heavy chain. The half-maximal concentration of the heptapeptide which induced an increase in the light-scattering intensity at 25 degrees C was about 110 microM, which was in the range of the dissociation constant of this peptide with F-actin. The polymerization of G-actin to F-actin by binding of the heptapeptide was further demonstrated by ultracentrifugal separation, Pi liberation, and electron microscopy. The polymerization of G-actin was induced only by the heptapeptide, but not by fragments of the heptapeptide. The well known acceleration of polymerization of G-actin by the myosin head may be due to the binding of G-actin with the S-site on the myosin head.  相似文献   

17.
A membrane fraction (MF2) has been purified from isolated microvilli of the MAT-C1 subline of the 13762 rat mammary ascites adenocarcinoma under conditions which cause F-actin depolymerization. This membrane preparation contains actin as a major component, although no filamentous structures are observed by transmission electron microscopy. Membranes were extracted with a Triton X-100-containing actin-stabilizing buffer (S buffer) or actin-destabilizing buffer (D buffer). In D buffer greater than 90% of metabolically labeled protein and glycoprotein was extracted, and 80-90% of these labeled species was extracted in S buffer. When S buffer extracts of MF2 were fractionated by either gel filtration on Sepharose 6 B or rate-zonal sucrose density gradient centrifugation, most of the actin was found to be intermediate in size between G- and F-actin. In D buffer most of the MF2 actin behaved as G-actin. Extraction and gel filtration of intact microvilli in S buffer also showed the presence of the intermediate form of actin, indicating that it did not arise during membrane preparation. When [35S]methionine-labeled G-actin from ascites cells was added to S buffer extracts of MF2 and chromatographed, all of the radioactivity chromatographed as G-actin, indicating that the intermediate form of actin did not result from an association of G-actin molecules during extraction or chromatography. The results of this study suggest that the microvillar membrane fraction is enriched in an intermediate form of actin smaller than F-actin and larger than G-actin.  相似文献   

18.
Ca2+-calmodulin-dependent polymerization of actin by myelin basic protein   总被引:1,自引:0,他引:1  
The interaction between myelin basic protein (MBP) and G-actin was studied under nonpolymerizing conditions, i.e.,2mM HEPES, pH 7.5, 0.1 mM CaCl2 and 0.2 mM ATP. Fluorescence studies using pyrenyl-actin and the measurements of ATP hydrolysis rate show that MBP induces changes in the structure of the actin monomer similar to those occurring during polymerization by salt. Electron microscope observations of the MBP-G-actin complex reveal the presence of filamentous structures which appear as separate filaments or as bundles of filaments in lateral association. These filaments are polar as visualized by attachment of heavy meromyosin. The biochemical data together with electron microscope observations suggest that the binding of MBP to G-actin under non-polymerizing conditions induces an interaction between actin monomers leading to the formation of filamentous structures which may be similar to F-actin filaments. The effects of MBP on G-actin can be reversed by calmodulin in the presence of Ca2+.  相似文献   

19.
The regulation of binding of G-actin to cytoplasmic domains of cell surface receptors is a common mechanism to control diverse biological processes. To model the regulation of G-actin binding to a cell surface receptor we used the cell-cell adhesion molecule carcinoembryonic antigen-related cell adhesion molecule 1 (CEACAM1-S) in which G-actin binds to its short cytoplasmic domain (12 amino acids; Chen, C. J., Kirshner, J., Sherman, M. A., Hu, W., Nguyen, T., and Shively, J. E. (2007) J. Biol. Chem. 282, 5749-5760). A liposome model system demonstrates that G-actin binds to the cytosolic domain peptide of CEACAM1-S in the presence of negatively charged palmitoyl-oleoyl phosphatidylserine (POPS) liposomes and Ca(2+). In contrast, no binding of G-actin was observed in palmitoyl-oleoyl phosphatidylcholine (POPC) liposomes or when a key residue in the peptide, Phe-454, is replaced with Ala. Molecular Dynamics simulations on CEACAM1-S in an asymmetric phospholipid bilayer show migration of Ca(2+) ions to the lipid leaflet containing POPS and reveal two conformations for Phe-454 explaining the reversible availability of this residue for G-actin binding. NMR transverse relaxation optimized spectroscopic analysis of (13)C-labeled Phe-454 CEACAM1-S peptide in liposomes plus actin further confirmed the existence of two peptide conformers and the Ca(2+) dependence of actin binding. These findings explain how a receptor with a short cytoplasmic domain can recruit a cytosolic protein in a phospholipid and Ca(2+)-specific manner. In addition, this model system provides a powerful approach that can be applied to study other membrane protein interactions with their cytosolic targets.  相似文献   

20.
He X  Liu YM  Wang W  Li Y 《Annals of botany》2006,98(1):49-55
BACKGROUND AND AIMS: Actin distribution in root hair tips is a controversial topic. Although the relationship between Ca2+ gradient and actin dynamics in plant tip-growth has been a focus of study, there is still little direct evidence on the exact relationship in root hair tip-growth. METHODS: G-actin was labelled by fluorescein isothiocyanate-DNase I. F-actin was labelled by tetramethylrhodamine isothiocyanate-phalloidin. Actin in root hairs of Triticum aestivum (wheat) was investigated using confocal laser-scanning microscopy. KEY RESULTS: Thick F-actin bundles did not extend into a region of approx. 5-10 microm from the tip of the growing root hairs, although they gave off branches of fine actin filaments in the hair tips. A tip-focused G-actin gradient was shown at the extreme apex of growing root hairs. In full-grown wheat root hairs, the tip-focused G-actin gradient disappeared while the thick F-actin bundles extended into the tips. BAPTA-AM, a Ca2+ disruption agent, also caused the tip-focused G-actin gradient to disappear and the diffuse F-actin bundles to appear in the tips of wheat root hairs. CONCLUSIONS: These results suggest that the tip-focused gradient of intracellular G-actin concentration at the extreme apex may be essential for root hair growth, and that preserving the tip-focused gradient needs a high Ca2+ concentration in the root hair tips.  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