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 172 毫秒
1.
Actin cleaved by the protease from Escherichia coli A2 strain between Gly42 and Val43 (ECP-actin) is no longer polymerizable when it contains Ca2+ as a tightly bound cation, but polymerizes when Mg2+ is bound. We have investigated the interactions of gelsolin with this actin with regard to conformational changes in the actin molecule induced by the binding of gelsolin. ECP-(Ca)actin interacts with gelsolin in a manner similar to that in which it reacts with intact actin, and forms a stoichiometric 2:1 complex. Despite the nonpolymerizability of ECP-(Ca)actin, this complex can act as a nucleus for the polymerization of intact actin, thus indicating that upon interaction with gelsolin, ECP-(Ca)actin undergoes a conformational change that enables its interaction with another actin monomer. By gel filtration and fluorometry it was shown that the binding of at least one of the ECP-cleaved actins to gelsolin is considerably weaker than of intact actin, suggesting that conformational changes in subdomain 2 of actin monomer may directly or allosterically affect actin-gelsolin interactions. On the other hand, interaction with gelsolin changes the conformation of actin within the DNase I-binding loop, as indicated by inhibition of limited proteolysis of actin by ECP and subtilisin. Cross-linking experiments with gelsolin-nucleated actin filaments using N,N-phenylene-bismaleimide (which cross-links adjacent actin monomers between Cys374 and Lys191) reveal that gelsolin causes a significant increase in the yield of the 115-kDa cross-linking product, confirming the evidence that gelsolin stabilizes or changes the conformation of the C-terminal region of the actin molecule, and these changes are propagated from the capped end along the filament. These results allow us to conclude that nucleation of actin polymerization by gelsolin is promoted by conformational changes within subdomain 2 and at the C-terminus of the actin monomer.  相似文献   

2.
Actin was purified from rat sarcoma-45 by using affinity chromatography on DNase I agarose. Actin was detected in the soluble and cytoskeletal fractions. The molecular mass of the protein was found to be equal to 45 kDa. The tumour actin specifically reacted with the antibody against skeletal muscle actin, inhibited the DNAase I activity and activated in the fibrillar state Mg(2+)-ATPases of sarcoma-45 and skeletal muscle myosins. The activating effect of the tumour protein was lower than that of its skeletal muscle counterpart. V8-protease peptide mapping revealed a similarity between tumour and brain actins. Sarcoma-45 actin was found to contain beta- and gamma-actin isoforms and an unusual isoform which appeared to be more acidic than the alpha-actin isoform.  相似文献   

3.
Dynamic properties of F-actin structure prompted suggestions (Squire, J. M., and Morris, E. P. (1998) FASEB J. 12, 761-771) that actin subdomain 2 movements play a role in thin-filament regulation. Using fluorescently labeled yeast actin mutants Q41C, Q41C/C374S, and D51C/C374S and azidonitrophenyl putrescine (ANP) Gln(41)-labeled alpha-actin, we monitored regulation-linked changes in subdomain 2. These actins had fully regulated acto-S1 ATPase activities, and emission spectra of regulated Q41C(AEDANS)/C374S and D51C(AEDANS)/C374S filaments did not reveal any calcium-dependent changes. Fluorescence energy transfer in these F-actins mostly occurred from Trp(340) and Trp(356) to 5-(2((acetyl)amino)ethyl)amino-naphthalene-1-sulfonate (AEDANS)-labeled Cys(41) or Cys(51) of adjacent same strand protomers. Our results show that fluorescence energy transfer between these residues is similar in the mostly blocked (-Ca(2+)) and closed (+Ca(2+)) states. Ca(2+) also had no effect on the excimer band in the pyrene-labeled Q41C-regulated actin, indicating virtually no change in the overlap of pyrenes on Cys(41) and Cys(374). ANP quenching of rhodamine phalloidin fluorescence showed that neither Ca(2+) nor S1 binding to regulated alpha-actin affects the phalloidin-probe distance. Taken together, our results indicate that transitions between the blocked, closed, and open regulatory states involve no significant subdomain 2 movements, and, since the cross-linked alpha-actin remains fully regulated, that subdomain 2 motions are not essential for actin regulation.  相似文献   

4.
The polymerization of scallop β-like actin is significantly slower than that of skeletal muscle α-actin. To reveal which steps of polymerization contribute to this difference, we estimated the efficiency of nucleation of the two actins, the rates of filament elongation at spontaneous and gelsolin-nucleated polymerization and the turnover rates of the filament subunits at steady-state. Scallop actin nucleated nearly twice less efficient than rabbit actin. In actin filaments with free ends, when dynamics at the barbed ends overrides that at the pointed ends, the relative association rate constants of α- and β-actin were similar, whereas the relative dissociation rate constant of β-ATP-actin subunits was 2- to 3-fold higher than that of α-actin. The 2- to 3-fold faster polymerization of skeletal muscle versus scallop Ca-actin was preserved with gelsolin-capped actin filaments when only polymerization at the pointed end is possible. With gelsolin-induced polymerization, the rate constants of dissociation of ATP-actin subunits from the pointed ends were similar, while the association rate constant of β-actin to the pointed filament ends was twice lower than that of α-actin. This difference may be of physiological relevance for functional intracellular sorting of actin isoforms.  相似文献   

5.
To assess more systematically functional differences among non-muscle and muscle actins and the effect of specific mutations on their function, we compared actin from Dictyostelium discoideum (D-actin) with actin from rabbit skeletal muscle (R-actin) with respect to the formation of filaments, their three-dimensional structure and mechanical properties. With Mg(2+) occupying the single high-affinity divalent cation-binding site, the course of polymerization is very similar for the two types of actin. In contrast, when Ca(2+ )is bound, D-actin exhibits a significantly longer lag phase at the onset of polymerization than R-actin. Crossover spacing and helical screw angle of negatively stained filaments are similar for D and R-F-actin filaments, irrespective of the tightly bound divalent cation. However, three-dimensional helical reconstructions reveal that the intersubunit contacts along the two long-pitch helical strands of D-(Ca)F-actin filaments are more tenuous compared to those in R-(Ca)F-actin filaments. D-(Mg)F-actin filaments on the other hand exhibit more massive contacts between the two long-pitch helical strands than R-(Mg)F-actin filaments. Moreover, in contrast to the structure of R-F-actin filaments which is not significantly modulated by the divalent cation, the intersubunit contacts both along and between the two long-pitch helical strands are weaker in D-(Ca)F-actin compared to D-(Mg)F-actin filaments. Consistent with these structural differences, D-(Ca)F-actin filaments were significantly more flexible than D-(Mg)F-actin.Taken together, this work documents that despite being highly conserved, muscle and non-muscle actins exhibit subtle differences in terms of their polymerization behavior, and the three-dimensional structure and mechanical properties of their F-actin filaments which, in turn, may account for their functional diversity.  相似文献   

6.
A myeloid leukemia cell line, M1, differentiates to macrophage and gains locomotive and phagocytic activity when incubated with conditioned medium (CM) from a fibroblast culture and bacterial endotoxin. To characterize the actin molecules before and after differentiation, the actin was purified through three sequential steps: DEAE-sephadex A- 50, polymerization/depolymerization, and sephadex G-150 chromatography. There were no essential differences between the inhibitory activity of actins from control M1 cells and CM-treated M1 cells on both DNase I and heavy meromyosin (HMMM) K(+)-EDTA-ATPase; the same dose response as with skeletal muscle actin took place. After the treatment with CM, however, the specific activity for the activation of HMMM Mg(2+)- ATPase by actin became two-fold that of untreated M1 actin, which was one third of the value for skeletal muscle actin. The V(max) for the control and the CM-treated M1 cell, as well as the skeletal muscle actins, proved to be the same. By contrast, the K(app) values for the control and CM-treated M1-cell actins were 3- and 1.5-fold the value for skeletal-muscle actin. This means that CM treatment of the M1 actin produced a twofold affinity for the Mg(2+)-ATPase of skeletal-muscle myosin. The critical concentrations for polymerization were compared under different salt concentrations and temperatures. Although no marked difference was found for the presence of 2 mM MgCl(2), 0.1 M KCl in place of MgCl(2) at 5 degrees C gave the following values: 0.1 mg/ml for skeletal-muscle actin, 0.7 mg/ml for control M1 actin, 0,5 mg/ml for CM- treated M1 actin, and 1.0 mg/ml for the D(-) subline that is insensitive to CM. Although the critical concentration of D(-) actin is extraordinarily high, this actin showed normal polymerization above the critical concentration. This together with the data presented in our previous paper, that the D(-) actin in the crude extract did not polymerize, suggests that an inhibitor for actin polymerization is present in the subline. The kinetics experiment at 0.1 M KCl and 25 degrees C revealed a slower polymerization of untreated M1- and D(-)-cell actins as compared with CM-treated M1 actin. This delayed polymerization was due to a delay during the nucleation stage, not during the elongation stage. By isoelectric focusing, the ratios of β- to γ-actin showed a marked difference depending on the states of cells: about 4.9 for control M1, 2.8 for CM-treated M1, and 7.6 for D(-)-subline actins. Tryptic peptide maps also revealed the presence of different peptides. Thus, the functional differences of actin before and after the differentiation was accompanied by some chemical changes in actin molecules.  相似文献   

7.
Vertebrate striated muscle contraction is regulated in a Ca(2+)-dependent fashion by tropomyosin (Tm) and troponin (Tn). This regulation involves shifts in the position of Tm and Tn on actin filaments and may include conformational changes in actin that are then communicated to myosin subfragment 1 (S1). To determine whether subdomain 2 of actin plays a role in this regulation, the DNase-I loop 38-52 of this subdomain was cleaved by subtilisin between residues Met(47) and Gly(48). Despite impaired unregulated function, the potentiation and regulation of cleaved actin movement in the in vitro motility assay was not significantly different from that of uncleaved actin. Stopped-flow measurements of ADP release from regulated and unregulated cleaved acto-S1 showed a marked increase in ADP release from acto-S1 in the presence of the regulatory complex. The enhancement of the actin affinity for S1 in the presence of regulatory proteins was greater for uncleaved than for cleaved F-actin. Finally, both cleaved and uncleaved actins protect myosin loop 1 from papain cleavage equally well. Our results suggest that the potentiation of actin function in the in vitro motility assay by regulatory proteins stems from changes in cross-bridge cycle kinetics. In addition, the unimpaired calcium-sensitive regulation of cleaved actin indicates that subdomain 2 conformation does not play an essential role in the regulation process.  相似文献   

8.
Muscle and yeast actins display distinct behavioral characteristics. To better understand the allosteric interactions that regulate actin function, we created a muscle/yeast hybrid actin containing a muscle-specific outer domain (subdomains 1 and 2) and a yeast inner domain (subdomains 3 and 4). Actin with muscle subdomain 1 and the two yeast N-terminal negative charges supported viability. The four negative charge muscle N terminus in a muscle subdomain 1 background caused death, but in the same background actin with three N-terminal acidic residues (3Ac/Sub1) led to sick but viable cells. Addition of three muscle subdomain 2 residues (3Ac/Sub12) produced no further deleterious effects. These hybrid actins caused depolarized cytoskeletons, abnormal vacuoles, and mitochondrial and endocytosis defects. 3Ac/Sub1 G-actin exchanged bound epsilonATP more slowly than wild type actin, and the exchange rate for 3Ac/Sub12 was even slower, similar to that for muscle actin. The mutant actins polymerized faster and produced less stable and shorter filaments than yeast actin, the opposite of that expected for muscle actin. Unlike wild type actin, in the absence of unbound ATP, polymerization led to ADP-F-actin, which rapidly depolymerized. Like yeast actin, the hybrid actins activated muscle myosin S1 ATPase activity only about one-eighth as well as muscle actin, despite having essentially a muscle actin-specific myosin-binding site. Finally, the hybrid actins behaved abnormally in a yeast Arp2/3-dependent polymerization assay. Our results demonstrate a unique sensitivity of yeast to actin N-terminal negative charge density. They also provide insight into the role of each domain in the control of the various functions of actin.  相似文献   

9.
Troponin C (TnC) is the Ca(2+)-binding regulatory protein of the troponin complex in muscle tissue. Vertebrate fast skeletal muscle TnCs bind four Ca(2+), while Akazara scallop (Chlamys nipponensis akazara) striated adductor muscle TnC binds only one Ca(2+) at site IV, because all the other EF-hand motifs are short of critical residues for the coordination of Ca(2+). Fourier transform infrared (FTIR) spectroscopy was applied to study coordination structure of Mg(2+) bound in a mutant Akazara scallop TnC (E142Q) in D(2)O solution. The result showed that the side-chain COO(-) groups of Asp 131 and Asp 133 in the Ca(2+)-binding site of E142Q bind to Mg(2+) in the pseudo-bridging mode. Mg(2+) titration experiments for E142Q and the wild-type of Akazara scallop TnC were performed by monitoring the band at about 1600 cm(-1), which is due to the pseudo-bridging Asp COO(-) groups. As a result, the binding constants of them for Mg(2+) were the same value (about 6 mM). Therefore, it was concluded that the side-chain COO(-) group of Glu 142 of the wild type has no relation to the Mg(2+) ligation. The effect of Mg(2+) binding in E142Q was also investigated by CD and fluorescence spectroscopy. The on-off mechanism of the activation of Akazara scallop TnC is discussed on the basis of the coordination structures of Mg(2+) as well as Ca(2+).  相似文献   

10.
Guan JQ  Almo SC  Reisler E  Chance MR 《Biochemistry》2003,42(41):11992-12000
The solution structures of isolated monomeric actins in their Mg(2+)-ATP and Ca(2+)-ATP bound forms and in complexes with gelsolin segment-1 have been probed using hydroxyl radicals (*OH) generated by synchrotron X-ray radiolysis. Proteolysis and mass spectrometry analysis of 28 peptides containing 58 distinct reactive probe sites within actin were used to monitor conformational variations linked to divalent cation and gelsolin segment-1 binding. The solvent accessibilities of the probe sites, as measured by footprinting in solution for the Ca(2+)-G-actin and Mg(2+)-G-actin complexes with gelsolin segment-1, were consistent with available crystallographic data. This included a specific protection at the contact interface between the partners, as revealed by reduced reactivity of peptide 337-359 in the complex. Aside from the specific protection indicated previously, the oxidation rates for the reactive residues of the isolated Ca(2+)-G-actin were similar to those of the actin gelsolin segment-1 complexes; however, the reactivity of numerous residues in the isolated Mg(2+)-G-actin form was significantly reduced. Specifically, Mg(2+)-G-actin has a set of protected sites relative to Ca(2+)-G-actin that suggest a structural reorganization in subdomains 4 and 2 and a C-terminus more closely packed onto subdomain 1. These conformational variations for isolated Mg(2+)-G-actin provide a structural basis for its greater tendency to polymerize into filaments as compared to Ca(2+)-G-actin.  相似文献   

11.
Various lines of evidence suggest that communication between tropomyosin and myosin in the regulation of vertebrate-striated muscle contraction involves yet unknown changes in actin conformation. Possible participation of loop 38-52 in this communication has recently been questioned based on unimpaired Ca(2+) regulation of myosin interaction, in the presence of the tropomyosin-troponin complex, with actin cleaved by subtilisin between Met(47) and Gly(48). We have compared the effects of actin cleavage by subtilisin and by protease ECP32, between Gly(42) and Val(43), on its interaction with myosin S1 in the presence and absence of tropomyosin or tropomyosin-troponin. Both individual modifications reduced activation of S1 ATPase by actin to a similar extent. The effect of ECP cleavage, but not of subtilisin cleavage, was partially reversed by stabilization of interprotomer contacts with phalloidin, indicating different pathways of signal transmission from the N- and C-terminal parts of loop 38-52 to myosin binding sites. ECP cleavage diminished the affinity to tropomyosin and reduced its inhibition of acto-S1 ATPase at low S1 concentrations, but increased the tropomyosin-mediated cooperative enhancement of the ATPase by S1 binding to actin. These effects were reversed by phalloidin. Subtilisin-cleaved actin more closely resembled unmodified actin than the ECP-modified actin. Limited proteolysis of the modified and unmodified F-actins revealed an allosteric effect of ECP cleavage on the conformation of the actin subdomain 4 region that is presumably involved in tropomyosin binding. Our results point to a possible role of the N-terminal part of loop 38-52 of actin in communication between tropomyosin and myosin through changes in actin structure.  相似文献   

12.
The existence of a unique sarcomeric actin is demonstrated in teleosts that possess substantial amounts of slow skeletal muscle in the trunk. The slow skeletal isotype is conserved. There is one amino acid substitution between Atlantic herring slow skeletal actin and the equivalent in salmonids. Conversely, the intra-species variation is considerable; 13 substitutions between different herring skeletal isotypes (slow versus fast). The isomorphisms (non-conservative underlined: residues, 2, 3, 103, 155, 160, 165, 278, 281, 310, 329, 358, 360 and 363) are restricted to sub-domains 1 and 3 and include the substitution Asp-360 in 'slow' to Gln in 'fast' which results in an electrophoretic shift at alkaline pH. The musculature of the trunk facilitates the preparation of isoactins for biochemical study. Herring slow skeletal G-actin (Ca.ATP) is more susceptible to thermal, and urea, -induced denaturation and subtilisin cleavage than that in fast skeletal, but more stable than the counterpart in salmonids (one substitution, Gln354Ala) highlighting the critical nature of actin's carboxyl-terminal insert. Fluorescent spectra of G-actin isoforms containing the isomorphism Ser155Ala in complexation with 2'-deoxy 3' O-(N'-Methylanthraniloyl) ATP infer similar polarity of the nucleotide binding cleft. An electrophoretic survey detected two skeletal actins in some (smelt and mackerel) but not all teleosts. One skeletal muscle actin was detected in frog and bird.  相似文献   

13.
This paper compares wild-type and two mutant beta-actins, one in which Ser14 was replaced by a cysteine, and a second in which both Ser14 and Asp157 were exchanged (Ser14-->Cys and Ser14-->Cys, Asp157-->Ala, respectively). Both of these residues are part of invariant sequences in the loops, which bind the ATP phosphates, in the interdomain cleft of actin. The increased nucleotide exchange rate, and the decreased thermal stability and affinity for DNase I seen with the mutant actins indicated that the mutations disturbed the interdomain coupling. Despite this, the two mutant actins retained their ATPase activity. In fact, the mutated actins expressed a significant ATPase activity even in the presence of Ca2+ ions, conditions under which actin normally has a very low ATPase activity. In the presence of Mg2+ ions, the ATPase activity of actin was decreased slightly by the mutations. The mutant actins polymerized as the wild-type protein in the presence of Mg2+ ions, but slower than the wild-type in a K+/Ca2+ milieu. Profilin affected the lag phases and elongation rates during polymerization of the mutant and wild-type actins to the same extent, whereas at steady-state, the concentration of unpolymerized mutant actin appeared to be elevated. Decoration of mutant actin filaments with myosin subfragment 1 appeared to be normal, as did their movement in the low-load motility assay system. Our results show that Ser14 and Asp157 are key residues for interdomain communication, and that hydroxyl and carboxyl groups in positions 14 and 157, respectively, are not necessary for ATP hydrolysis in actin.  相似文献   

14.
Saeki K  Wakabayashi T 《Biochemistry》2000,39(6):1324-1329
To probe the mechanism by which Ca(2+) activates muscle contraction through tropomyosin and troponin, we have produced mutant actins using Dictyostelium discoideum. We focused on the sequence 228-232 (QTAAS) that is located in subdomain 4 of actin, because the chimera actin in which this sequence was replaced by KAYKE showed not only poorer tropomyosin binding but also the unexpected "higher Ca(2+) activation" [Saeki, K., et al. (1996) Biochemistry 35, 14465-14472]. We found that this higher Ca(2+) activation is solely due to the A230Y mutagenesis. Because A230Y mutant actin showed normal tropomyosin binding, the higher Ca(2+) activation is not the consequence of poorer tropomyosin binding. The significance of these results is discussed in view of a three-state model [McKillop, D. F., Geeves, M. A. (1993) Biophys. J. 65, 693-701].  相似文献   

15.
Crude actin extracts from acetone-dried powder of the muscle layer of bovine aorta contain an actin-modulating protein which promotes nucleation of actin monomers and decreases the average length of actin filaments in a Ca2+-dependent manner. This observation has allowed the development of an improved purification procedure for aorta actin which increases the yield 2- to 3-times. The actin obtained with this procedure consists of 77% alpha- and 23% gamma-isoelectric species. Pure aorta actin is indistinguishable from actins from skeletal, cardiac and chicken-gizzard smooth muscle in its polymerization rate, critical concentration, and reduced viscosity when polymerized with KCl at 25 degrees C. It differs from sarcomeric actins, but not from chicken-gizzard smooth muscle actin, in the temperature dependence of polymerization equilibria in KCl. This difference correlates with the amino acid replacements Val-17----Cys-17 and Thr-89----Ser-89, supporting a conclusion drawn from other studies that the N-terminal portion of actin polypeptide chain contains sites important for polymerization.  相似文献   

16.
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   

17.
Leishmania actin was cloned, overexpressed in baculovirus-insect cell system, and purified to homogeneity. The purified protein polymerized optimally in the presence of Mg2+ and ATP, but differed from conventional actins in its following properties: (i) it did not polymerize in the presence of Mg2+ alone, (ii) it polymerized in a restricted range of pH 7.0-8.5, (iii) its critical concentration for polymerization was found to be 3-4-fold lower than of muscle actin, (iv) it predominantly formed bundles rather than single filaments at pH 8.0, (v) it displayed considerably higher ATPase activity during polymerization, (vi) it did not inhibit DNase-I activity, and (vii) it did not bind the F-actin-binding toxin phalloidin or the actin polymerization disrupting agent Latrunculin B. Computational and molecular modeling studies revealed that the observed unconventional behavior of Leishmania actin is related to the diverged amino acid stretches in its sequence, which may lead to changes in the overall charge distribution on its solvent-exposed surface, ATP binding cleft, Mg2+ binding sites, and the hydrophobic loop that is involved in monomer-monomer interactions. Phylogenetically, it is related to ciliate actins, but to the best of our knowledge, no other actin with such unconventional properties has been reported to date. It is therefore suggested that actin in Leishmania may serve as a novel target for design of new antileishmanial drugs.  相似文献   

18.
We cloned, from a cDNA library, an alpha-actin sequence from a salamander (Pleurodeles waltlii), which codes for the 125 COOH-terminal amino acid residues of a skeletal muscle actin (without any difference from the corresponding protein of warm blood vertebrates). An important conservation in the 3' untranslated region between this sequence and skeletal alpha-actin genes of chicken and man was noted. These results demonstrate, contrary to what was thought previously, that there exists in salamander a true skeletal alpha-actin gene. The results suggest that striated muscle actin genes in lower vertebrates could be a mosaic of cardiac and skeletal-specific amino acid residues, and that the divergence between these two types of genes is older than the NH2-terminal analysis of actins suggested previously.  相似文献   

19.
A simple and effective technique to produce monospecific polyclonal antibodies of high affinity against actin is described. In this procedure, rabbit skeletal muscle actin in the 1:1 complex with bovine pancreatic deoxyribonuclease I is used as antigen to immunize rabbits. The antisera obtained are shown to contain antibodies against both actin and deoxyribonuclease I. By affinity chromatography the two antibody preparations were separated and characterized. The affinity-purified anti-deoxyribonuclease I and anti-actin do not show cross-reactivity. Thus, anti-deoxyribonuclease I inhibits the enzymic activity of deoxyribonuclease I and stains the enzyme after Western blotting. Affinity-purified anti-actin does not inhibit deoxyribonuclease I activity and stains only actin after Western blotting. The affinity-purified anti-actin can be used in a number of different actin-detecting techniques such as in immunohistochemistry and in immunoblotting techniques. This antibody recognizes only actins from muscular tissues with high affinity. Immunoblots of polyacrylamide gels in the presence of ampholytes (IEF) indicate that this antibody only recognizes the alpha-variants of actin. Thus, the skeletal and cardiac alpha-actins are recognized but not the smooth muscle gamma-isoform and the cytoplasmic actins. Vascular smooth muscle alpha-actin is not recognized when using immunoblotting or enzyme-linked immunosorbent techniques. On frozen sections, however, the anti-actin antibody clearly stained vascular smooth muscle cells. Epitope analysis using actin fragments generated by limited proteolysis and selective cleavage using hydroxylamine indicate that this antibody is directed against a rather limited region within the N-terminus of actin.  相似文献   

20.
Calcium ions produce a 3-4-fold stimulation of the actin-activated ATPase activities of phosphorylated myosin from bovine pulmonary artery or chicken gizzard at 37 degrees C and at physiological ionic strengths, 0.12-0.16 M. Actins from either chicken gizzard or rabbit skeletal muscle stimulate the activity of phosphorylated myosin in a Ca2+-dependent manner, indicating that the Ca2+ sensitivity involves myosin or a protein associated with it. Partial loss of Ca2+ sensitivity upon treatment of phosphorylated gizzard myosin with low concentrations of chymotrypsin and the lack of any change on similar treatment of actin supports the above conclusion. Although both actins enhance ATPase activity, activation by gizzard actin exhibits Ca2+ dependence at higher temperatures or lower ionic strengths than does activation by skeletal muscle actin. The Ca2+ dependence of the activity of phosphorylated heavy meromyosin is about half that of myosin and is affected differently by temperature, ionic strength and Mg2+, being independent of temperature and optimal at lower concentrations of NaCl. Raising the concentration of Mg2+ above 2-3 mM inhibits the activity of heavy meromyosin but stimulates that of myosin, indicating that Mg2+ and Ca2+ activate myosin at different binding sites.  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