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.
Salmonids utilize a unique, class II isoactin in slow skeletal muscle. This actin contains 12 replacements when compared with those from salmonid fast skeletal muscle, salmonid cardiac muscle and rabbit skeletal muscle. Substitutions are confined to subdomains 1 and 3, and most occur after residue 100. Depending on the pairing, the 'fast', 'cardiac' and rabbit actins share four, or fewer, substitutions. The two salmonid skeletal actins differ nonconservatively at six positions, residues 103, 155, 278, 281, 310 and 360, the latter involving a change in charge. The heterogeneity has altered the biochemical properties of the molecule. Slow skeletal muscle actin can be distinguished on the basis of mass, hydroxylamine cleavage and electrophoretic mobility at alkaline pH in the presence of 8 m urea. Further, compared with its counterpart in fast muscle, slow muscle actin displays lower activation of myosin in the presence of regulatory proteins, and weakened affinity for nucleotide. It is also less resistant to urea- and heat-induced denaturation. The midpoints of the change in far-UV ellipticity of G-actin versus temperature are approximately 45 degrees C ('slow' actin) and approximately 56 degrees C ('fast' actin). Similar melting temperatures are observed when thermal unfolding is monitored in the aromatic region, and is suggestive of differential stability within subdomain 1. The changes in nucleotide affinity and stability correlate with substitutions at the nucleotide binding cleft (residue 155), and in the C-terminal region, two parts of actin which are allosterically coupled. Actin is concluded to be a source of skeletal muscle plasticity.  相似文献   

2.
C Frieden  K Patane 《Biochemistry》1985,24(15):4192-4196
The role of adenosine 5'-triphosphate (ATP) in the Mg2+-induced conformational change of rabbit skeletal muscle G-actin has been investigated by comparing actin containing bound ADP with actin containing bound ATP. As previously described [Frieden, C. (1982) J. Biol. Chem. 257, 2882-2886], N-acetyl-N'-(5-sulfo-1-naphthyl)ethylenediamine-labeled G-actin containing ATP undergoes a time-dependent Mg2+-induced fluorescence change that reflects a conformational change in the actin. Addition of Mg2+ to labeled G-actin containing ADP gives no fluorescence change, suggesting that the conformational change does not occur. The fluorescence change can be restored on the addition of ATP. Examination of the time courses of these experiments suggests that ATP must replace ADP prior to the Mg2+-induced change. The Mg2+-induced polymerization of actin containing ADP is extraordinarily slow compared to that of actin containing ATP. The lack of the Mg2+-induced conformational change, which is an essential step in the Mg2+-induced polymerization, is probably the cause for the very slow polymerization of actin containing ADP. On the other hand, at 20 degrees C, at pH 8, and in 2 mM Mg2+, the elongation rate from the slow growing end of an actin filament, measured by using the protein brevin to block growth at the fast growing end, is only 4 times slower for actin containing ADP than for actin containing ATP.  相似文献   

3.
Conformational changes in subdomain 2 of actin were investigated using fluorescence probes dansyl cadaverine (DC) or dansyl ethylenediamine (DED) covalently attached to Gln41. Examination of changes in the fluorescence emission spectra as a function of time during Ca2+/Mg2+ and ATP/ADP exchange at the high-affinity site for divalent cation-nucleotide complex in G-actin confirmed a profound influence of the type of nucleotide but failed to detect a significant cation-dependent difference in the environment of Gln41. No significant difference between Ca- and Mg-actin was also seen in the magnitude of the fluorescence changes resulting from the polymerization of these two actin forms. Evidence is presented that earlier reported cation-dependent differences in the conformation of the loop 38-52 may be related to time-dependent changes in the conformation of subdomain 2 in DED- or DC-labeled G-actin, accelerated by substitution of Mg2+ for Ca2+ in CaATP-G-actin and, in particular, by conversion of MgATP- into MgADP-G-actin. These spontaneous changes are associated with a denaturation-driven release of the bound nucleotide that is promoted by two effects of DED or DC labeling: lowered affinity of actin for nucleotide and acceleration of ATP hydrolysis on MgATP-G-actin that converts it into a less stable MgADP form. Evidence is presented that the changes in the environment of Gln41 accompanying actin polymerization result in part from the release of Pi after the hydrolysis of ATP on the polymer. A similarity of this change to that accompanying replacement of the bound ATP with ADP in G-actin is discussed.  相似文献   

4.
Structural effects of yeast cofilin on skeletal muscle and yeast actin were examined in solution. Cofilin binding to native actin was non-cooperative and saturated at a 1:1 molar ratio, with K(d)相似文献   

5.
Cofilin, a member of the actin-depolymerizing factor (ADF)/cofilin family of proteins, is a key regulator of actin dynamics. Cofilin binds to monomer (G-) and filamentous (F-) actin, severs the filaments, and increases their turnover rate. Electron microscopy studies suggested cofilin interactions with subdomains 2 and 1/3 on adjacent actin protomers in F-actin. To probe for the presence of a cryptic cofilin binding site in subdomain 2 in G-actin, we used transglutaminase-mediated cross-linking, which targets Gln41 in subdomain 2. The cross-linking proceeded with up to 85% efficiency with skeletal alpha-actin and WT yeast actin, yielding a single product corresponding to a 1:1 actin-cofilin complex but was strongly inhibited in Q41C yeast actin (in which Q41 was substituted with cysteine). LC-MS/MS analysis of the proteolytic fragments of this complex mapped the cross-linking to Gln41 on actin and Gly1 on recombinant yeast cofilin. The actin-cofilin (AC) heterodimer was purified on FPLC for analytical ultracentrifugation and electron microscopy analysis. Sedimentation equilibrium and velocity runs revealed oligomers of AC in G-actin buffer. In the presence of excess cofilin, the covalent AC heterodimer bound a second cofilin, forming a 2:1 cofilin/actin complex, as revealed by sedimentation results. Under polymerizing conditions the cross-linked AC formed mostly short filaments, which according to image reconstruction were similar to uncross-linked actin-cofilin filaments. Although a majority of the cross-linking occurs at Gln41, a small fraction of the AC cross-linked complex forms in the Q41C yeast actin mutant. This secondary cross-linking site was sequenced by MALDI-MS/MS as linking Gln360 in actin to Lys98 on cofilin. Overall, these results demonstrate that the region around Gln41 (subdomain 2) is involved in a weak binding of cofilin to G-actin.  相似文献   

6.
Partially purified smooth muscle (chicken gizzard) actomyosin contains two major substrates of cAMP-dependent protein kinase: a protein of Mr = 130,000, identified as the calmodulin-dependent myosin light chain kinase, and a protein of Mr = 42,000. This latter protein was shown by a variety of electrophoretic procedures to be actin. Purified smooth muscle actin also was phosphorylated by the catalytic subunit of cAMP-dependent protein kinase. The rate of phosphorylation of smooth muscle actin was significantly enhanced by depolyjerization of actin. A maximum of 2.0 mol phosphate could be incorporated per mol G-actin. Skeletal muscle F-actin was not significantly phosphorylated by protein kinase; however, skeletal G-actin is a substrate for the protein kinase although its rate of phosphorylation was significantly slower than that of smooth muscle G-actin.  相似文献   

7.
Complete amino acid sequences for four mammalian muscle actins are reported: bovine skeletal muscle actin, bovine cardiac actin, the major component of bovine aorta actin, and rabbit slow skeletal muscle actin. The number of different actins in a higher mammal for which full amino acid sequences are now available is therefore increased from two to five. Screening of different smooth muscle tissues revealed in addition to the aorta type actin a second smooth muscle actin, which appears very similar if not identical to chicken gizzard actin. Since the sequence of chicken gizzard actin is known, six different actins are presently characterized in a higher mammal.
The two smooth muscle actins—bovine aorta actin and chicken gizzard actin—differ by only three amino acid substitutions, all located in the amino-terminal end. In the rest of their sequences both smooth muscle actins share the same four amino acid substitutions, which distinguish them from skeletal muscle actin. Cardiac muscle actin differs from skeletal muscle actin by only four amino acid exchanges. No amino acid substitutions were found when actins from rabbit fast and slow skeletal muscle were compared.
In addition we summarize the amino acid substitution patterns of the six different mammalian actins and discuss their tissue specificity. The results show a very close relationship between the four muscle actins in comparison to the nonmuscle actins. The amino substitution patterns indicate that skeletal muscle actin is the highest differentiated actin form, whereas smooth muscle actins show a noticeably closer relation to nonmuscle actins. By these criteria cardiac muscle actin lies between skeletal muscle actin and smooth muscle actins.  相似文献   

8.
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   

9.
J F Hoh 《Biochemistry》1975,14(4):742-747
Mammalian nerves to fast and slow muscles have the remarkable property of changing the speed of contraction of muscles following cross-reinnervation. The biochemical basis of speed transformation is the change in myosin in ATPase activity. This paper provides electrophoretic evidence for structural changes in myosin from cross-reinnervated muscles. A method is described for the separation of intact fast and slow muscle myosins by polyacrylamide gel electrophoresis. This method utilizes the fact that ATP and its analogs prevent the formation of myosin polymers in low ionic strength buffers. In this system, normal fast muscle myosin has a higher electrophoretic mobility than slow muscle myosin. Normal rat soleus myosin has a major slow and a minor fast component due to two populations of muscle fibers. The same muscle cross-reinnervated by a fast muscle nerve shows only the fast component, The normal, homogeneous fast extensor digitorum longus muscle has only the electrophoretically fast myosin, but following cross-reinnervation it shows both fast and slow components. These results suggest that mammalian motor nerves can induce or suppress the expression of genes that code for fast and slow skeletal muscle myosins.  相似文献   

10.
We have utilized a key biochemical determinant of muscle fiber type, myosin isoform expression, to investigate the initial developmental program of future fast and slow skeletal muscle fibers. We examined myosin heavy chain (HC) phenotype from the onset of myogenesis in the limb bud muscle masses of the chick embryo through the differentiation of individual fast and slow muscle masses, as well as in newly formed myotubes generated in adult muscle by weight overload. Myosin HC isoform expression was analyzed by immunofluorescence localization with a battery of anti-myosin antibodies and by electrophoretic separation with SDS-PAGE. Results showed that the initial myosin phenotype in all skeletal muscle cells formed during the embryonic period (until at least 8 days in ovo) consisted of expression of a myosin HC which shares antigenic and electrophoretic migratory properties with ventricular myosin and a distinct myosin HC which shares antigenic and electrophoretic migratory properties with fast skeletal isomyosin. Similar results were observed in newly formed myotubes in adult muscle. Future fast and slow muscle fibers could only be discriminated from each other in developing limb bud muscles by the onset of expression of slow skeletal myosin HC at 6 days in ovo. Slow skeletal myosin HC was expressed only in myotubes which became slow fibers. These findings suggest that the initial commitment of skeletal muscle progenitor cells is to a common skeletal muscle lineage and that commitment to a fiber-specific lineage may not occur until after localization of myogenic cells in appropriate premuscle masses. Thus, the process of localization, or events which occur soon thereafter, may be involved in determining fiber type.  相似文献   

11.
Isolation and characterization of covalently cross-linked actin dimer   总被引:5,自引:0,他引:5  
Covalently cross-linked actin dimer was isolated from rabbit skeletal muscle F-actin reacted with phenylenebismaleimide (Knight, P., and Offer, G. (1978) Biochem. J. 175, 1023-1032). The UV spectrum of the purified cross-linked actin dimer, in a nonpolymerizing buffer, was very similar to that of native F-actin and not to the spectrum of G-actin. Cross-linked actin dimer polymerized to filaments that were indistinguishable in the electron microscope from F-actin made from native G-actin and that were similar to native F-actin in their ability to activate the Mg2+-ATPase of myosin subfragment-1. The critical concentrations of polymerization of cross-linked actin dimer in 0.5 mM and 2.0 mM MgCl2, 2 to 4 microM, and 1 to 2 microM, respectively, were similar to the values for native G-actin. Cross-linked actin dimer contained 2 mol of bound nucleotide/mol of dimer. One bound nucleotide exchanged with ATP in solution with a t 1/2 of 55 min and with ADP with a t 1/2 of 5 h. The second bound nucleotide exchanged much more slowly. The more rapidly exchangeable site contained 10 to 15% bound ADP.Pi and 85 to 90% bound ATP while the second site contained much less, if any, bound ADP.Pi. Cross-linked actin dimer had an ATPase activity in 0.5 mM MgCl2 that was 7 times greater than the ATPase activity of native G-actin and that was also stimulated by cytochalasin D. These data are discussed in relation to the possible role of ATP in actin polymerization and function with the speculation that the cross-linked actin dimer may serve simultaneously as a useful model for each of the two different ends of native F-actin.  相似文献   

12.
The actin concentration and monomer-polymer ratio in developing chicken skeletal muscle were determined by means of a DNase I inhibition assay. The concentration of G-actin in embryonic muscle was much higher than the critical concentration for polymerization of purified actin. As muscle development progressed, the amount of total actin remarkably increased, whereas the concentration of G-actin markedly decreased, and finally in adults reached the critical concentration for polymerization of purified actin. When the monomeric actin in the soluble fraction of embryonic muscle was purified, the critical concentration for polymerization of the embryonic actin decreased to the same value as that of adult skeletal muscle actin. On the other hand, there was no difference between the crude and purified actin in the type of actin. They consisted of alpha-, beta-, and gamma-actins; their amounts were in the order, beta greater than gamma greater than alpha. Furthermore, polymerization of the monomeric actin in the soluble fraction of embryonic muscle was induced by the addition of myosin or HMM. The large amount of monomeric actin in the embryonic skeletal muscle may be due to the presence of some factor(s) which inhibits actin polymerization and also to an insufficiency of myosin.  相似文献   

13.
Mechanism of K+-induced actin assembly   总被引:14,自引:6,他引:8       下载免费PDF全文
The assembly of highly purified actin from Dictyostelium discoideum amoebae and rabbit skeletal muscle by physiological concentrations of KCI proceeds through successive stages of (a) rapid formation of a distinct monomeric species referred to as KCI-monomer, (b) incorporation of KCI-monomers into an ATP-containing filament, and (c) ATP hydrolysis that occurs significantly after the incorporation event. KCI-monomer has a conformation which is distinct from that of either conventional G- or F-actin, as judged by UV spectroscopy at 210-220 nm and by changes in ATP affinity. ATP is not hydrolyzed during conversion of G-actin to KCI-monomer. KCI-monomer formation precedes filament formation and may be necessary for the assembly event. Although incorporation of KCI-monomers into filaments demonstrates lagphase kinetics by viscometry, both continuous absorbance monitoring at 232 nm and rapid sedimentation of filaments demonstrate hyperbolic assembly curves. ATP hydrolysis significantly lags the formation of actin filaments. When half of the actin has assembled, only 0.1 to 0.2 mole of ATP are hydrolyzed per mole of actin present as filaments.  相似文献   

14.
The regulation of actin is key for controlled cellular function. Filaments are regulated by actin-binding proteins, but the nucleotide state of actin is also an important factor. From extended molecular dynamics simulations, we find that both nucleotide states of the actin monomer have significantly less twist than their crystal structures and that the ATP monomer is flatter than the ADP form. We also find that the filament’s pointed end is flatter than the remainder of the filament and has a conformation distinct from G-actin, meaning that incoming monomers would need to undergo isomerization that would weaken the affinity and slow polymerization. Conversely, the barbed end of the filament takes on a conformation nearly identical to the ATP monomer, enhancing ATP G-actin’s ability to polymerize as compared with ADP G-actin. The thermodynamic penalty imposed by differences in isomerization for the ATP and ADP growth at the barbed end exactly matches experimental results.  相似文献   

15.
In the previous study (Podlubnaya et al., 1999, J. Struc. Biol. 127, 1-15) Ca2+-induced reversible structural transitions in synthetic filaments of pure fast skeletal and cardiac muscle myosins were observed under rigor conditions (-Ca2+/+Ca2+). In the present work these studies have been extended to new more order-producing conditions (presence of ATP in the absence of Ca2+) aimed at arresting the relaxed structure in synthetic filaments of both fast and slow skeletal muscle myosin. Filaments were formed from column-purified myosins (rabbit fast skeletal muscle and rabbit slow skeletal semimebranosusproprius muscle). In the presence of 0.1 mM free Ca2+, 3 mM Mg2+ and 2 mM ATP (activating conditions) these filaments had a spread structure with a random arrangement of myosin heads and subfragments 2 protruding from the filament backbone. Such a structure is indistinguishable from the filament structures observed previously for fast skeletal, cardiac (see reference cited above) and smooth (Podlubnaya et al., 1999, J. Muscle Res. Cell Motil. 20, 547-554) muscle myosins in the presence of 0.1 mM free Ca2+. In the absence of Ca2+ and in the presence of ATP (relaxing conditions) the filaments of both studied myosins revealed a compact ordered structure. The fast skeletal muscle myosin filaments exhibited an axial periodicity of about 14.5 nm and which was much more pronounced than under rigor conditions in the absence of Ca2+ (see the first reference cited). The slow skeletal muscle myosin filaments differ slightly in their appearance from those of fast muscle as they exhibit mainly an axial repeat of about 43 nm while the 14.5 nm repeat is visible only in some regions. This may be a result of a slightly different structural properties of slow skeletal muscle myosin. We conclude that, like other filaments of vertebrate myosins, slow skeletal muscle myosin filaments also undergo the Ca2+-induced structural order-disorder transitions. It is very likely that all vertebrate muscle myosins possess such a property.  相似文献   

16.
Circular dichroic spectra of native, EDTA-treated and heat-denatured G-actin from chicken gizzard smooth muscle are virtually the same as those of rabbit skeletal muscle actin. The rates of changes produced by EDTA or heat in the secondary structure are, however, higher in the case of gizzard actin. Similar differences were found in the rates of inactivation as measured by loss of polymerizability during incubation with EDTA or Dowex 50. The results are explicable in terms of local differences in the conformation at specific site(s) important for maintaining the native state of actin monomer. Involvement of the ATP binding site was shown by measuring the equilibrium constant for the binding of ATP to the two actins. Difference in the conformation of some additional site(s) is indicated by a higher rate constant of inactivation of nucleotide-free actin observed for gizzard actin. No significant difference was found in the equilibrium constant for the binding of Ca2+ at the single high-affinity site in gizzard and skeletal muscle actin. Comparison of inactivation kinetics of actin from chicken gizzard, rabbit skeletal, bovine aorta, and bovine cardiac muscle suggests that the amino acid replacements Val-17----Cys-17 and/or Thr-89----Ser-89 have a destabilizing effect on the native conformation of G-actin. The results indicate that deletion of the acidic residue at position 1 of the amino acid sequence has no effect on the conformation of the ATP binding site and the high-affinity site for divalent cation as well.  相似文献   

17.
Millisecond internal dynamics of native and inactivated actin from rabbit skeletal muscle was examined using room temperature phosphorescence. Inactivated actin was prepared by incubation of G-actin at 70 degrees C, by treatment with 4 M urea or 1.5 M guanidinium hydrochloride, renaturation from fully unfolded state or by Ca2+ ion removal. It was shown that inactivation of actin, irrespective of the denaturation procedure applied, leads to a sharp decrease of millisecond fluctuations of the protein structure. Restriction of the slow intramolecular mobility in inactivated actin can result from changes of the protein conformation and/or specific association of macromolecules.  相似文献   

18.
E Kim  M Motoki  K Seguro  A Muhlrad    E Reisler 《Biophysical journal》1995,69(5):2024-2032
Gln-41 on G-actin was specifically labeled with a fluorescent probe, dansyl ethylenediamine (DED), via transglutaminase reaction to explore the conformational changes in subdomain 2 of actin. Replacement of Ca2+ with Mg2+ and ATP with ADP on G-actin produced large changes in the emission properties of DED. These substitutions resulted in blue shifts in the wavelength of maximum emission and increases in DED fluorescence. Excitation of labeled actin at 295 nm revealed energy transfer from tryptophans to DED. Structure considerations and Cu2+ quenching experiments suggested that Trp-79 and/or Trp-86 serves as energy donors to DED. Energy transfer from these residues to DED on Gln-41 increased with the replacement of Ca2+ with Mg2+ and ATP with ADP. Polymerization of Mg-G-actin with MgCl2 resulted in much smaller changes in DED fluorescence than divalent cation substitution. This suggests that the conformation of loop 38-52 on actin is primed for the polymerization reaction by the substitution of Ca2+ with Mg2+ on G-actin.  相似文献   

19.
Mechanism for nucleotide exchange in monomeric actin   总被引:1,自引:0,他引:1  
C Frieden  K Patane 《Biochemistry》1988,27(10):3812-3820
Rabbit skeletal muscle G-actin has been treated to obtain ADP, 1,N6-ethenoadenosine diphosphate (epsilon-ADP), or 1,N6-ethenoadenosine triphosphate (epsilon-ATP) at the nucleotide binding site and either Mg2+ or Ca2+ at high- and moderate-affinity metal binding sites. Apparent rates or rate constants for the displacement of the actin-bound nucleotides by epsilon-ATP or ATP have been obtained by stopped-flow measurements at pH 8 and 20 degrees C of the fluorescence difference between bound and free epsilon-ATP or epsilon-ADP. In the presence of Ca2+, displacement of ADP by epsilon-ATP or epsilon-ADP by ATP is a biphasic process, but in the presence of low (less than 10 microM) Mg2+ concentrations, it is a slow first-order process. At high levels of Mg2+ (greater than 50 microM), low ADP concentrations displace epsilon-ATP from G-actin as a consequence of Mg2+ binding to moderate-affinity sites on the actin. Displacement of epsilon-ATP by ATP in the presence of either Ca2+ or Mg2+ is slow at low ATP concentrations, but the rate is increased by high ATP concentrations. Using ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, we find that nucleotide exchange is affected differently by the removal of Ca2+ from the high-affinity site compared to Ca2+ removal from moderate-affinity sites. A mechanism for the displacement reaction is proposed in which there are two forms of an actin-ADP complex and metal binding influences the ratio of these forms as well as the binding of ATP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Addition of low concentrations (0.2--2.0 mM) of EGTA to rabbit skeletal muscle G-actin in the presence of ATP caused increase in viscosity. The effect is probably due to chelation of Ca2+. EGTA-polymerized actin was sedimented in the ultracentrifuge as a pellet which could be depolymerized in the presence of Ca2+ and then repolymerized. Electron microscopy indicated that formation of filamentous actin which appears to be somewhat more flexible than F-actin obtained by polymerization with KCl. The EGTA-polymerized actin was dissociated by DNAase I faster than KCl-polymerized actin. F-Actin can thus be stable also in very low ionic strength media if Ca2+ is removed whereas for G-actin to be the only form of the protein in such media, micromolar concentrations of Ca2+ must be present.  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