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 740 毫秒
1.
To determine the reason for the inviability of Saccharomyces cerevisiae with skeletal muscle actin, we introduced into yeast actin the first variant muscle residue from the C-terminal end, H372R. Arg is also found at this position in non-yeast nonmuscle actins. The substitution caused retarded growth on glucose and an inability to use glycerol as a sole carbon source. The mitochondria were clumped and had lost their DNA, the vacuole appeared hypervesiculated, and the actin cytoskeleton became somewhat depolarized. Introduction of the second muscle actin-specific substitution, S365A, rescued these defects. Suppression was also achieved by introducing the four acidic N-terminal residues of muscle actin in place of the two found in yeast actin. The H372R substitution results in an increase in polymerization-dependent fluorescence of Cys-374 pyrene-labeled actin. H372R actin polymerizes slightly faster than wild-type (WT) actin. Yeast actin-related proteins 2 and 3 (Arp2/3) accelerates the polymerization of H372R actin to a much greater extent than WT actin. The two suppressors did not affect the rate of H372R actin polymerization in the absence of an Arp2/3 complex. In contrast, the S365A substitution dampened the rate of Arp2/3 complex-stimulated H372R actin polymerization, and the addition of the four acidic N-terminal residues caused this rate to decrease below that observed with WT actin in the presence of Arp2/3. Structural analysis of the mutations suggests the presence of stringent steric and ionic requirements for the bottom of actin subdomain 1 and also suggests that there is allosteric communication through subdomain 1 within the actin monomer between the N and C termini.  相似文献   

2.
Actin can exist in multiple conformations necessary for normal function. Actin isoforms, although highly conserved in sequence, exhibit different biochemical properties and cellular roles. We used amide proton hydrogen/deuterium (HD) exchange detected by mass spectrometry to analyze conformational differences between Saccharomyces cerevisiae and muscle actins in the G and F forms to gain insight into these differences. We also utilized HD exchange to study interdomain and allosteric communication in yeast-muscle hybrid actins to better understand the conformational dynamics of actin. Areas showing differences in HD exchange between G- and F-actins are areas of intermonomer contacts, consistent with the current filament models. Our results showed greater exchange for yeast G-actin compared with muscle actin in the barbed end pivot region and areas in subdomains 1 and 2 and for F-actin in monomer-monomer contact areas. These results suggest greater flexibility of the yeast actin monomer and filament compared with muscle actin. For hybrid G-actins, the muscle-like and yeastlike parts of the molecule generally showed exchange characteristics resembling their parent actins. A few exceptions were a peptide on top of subdomain 2 and the pivot region between subdomains 1 and 3 with muscle actin-like exchange characteristics although the areas were yeastlike. These results demonstrate that there is cross-talk between subdomains 1 and 2 and the large and small domains. Hybrid F-actin data showing greater exchange compared with both yeast and muscle actins are consistent with mismatched yeast-muscle interfaces resulting in decreased stability of the hybrid filament contacts.The ability of actin to engage in a wide range of physiological functions requires that it be subject to complex spatial and temporal control by a large array of actin-binding proteins (13). Such regulation demands that both the monomeric and filamentous forms of actin be able to exist in a number of different conformations. Some of these may be differentially recognized by different actin-binding proteins, and some might actually be induced by the interaction of actin with these proteins.Actin is highly conserved from yeast to humans. There is 87% sequence identity between yeast and muscle actin and 91% sequence identity between yeast and nonmuscle actins. Even with this high sequence similarity and minor differences in crystal structures, there are some major differences in yeast and muscle actin behavior. Yeast, compared with muscle actin, polymerizes faster and exchanges its bound nucleotide faster. In contrast to muscle actin, the yeast actin filament releases its Pi almost immediately after ATP hydrolysis and the filament fragments more readily (4). Yeast cells cannot survive with muscle actin as the sole actin, and with β-nonmuscle actin as the only actin, yeast cells are very sick (5). A set of biochemical data indicates that the muscle filament is less flexible than yeast (6, 7). However, conformational and structural differences that could explain these behavioral differences are unknown.The actin molecule is divided into two domains, the small domain consisting of subdomains 1 and 2 and the large domain consisting of subdomains 3 and 4. Modeling studies (8, 9) have suggested that each of the subdomains can move, and biochemical studies have suggested allosteric interactions between the subdomains. However, the nature of these interactions has not been elucidated on a molecular level.To try to gain insight into what parts of the actin molecule are important for the behavioral differences observed between yeast and muscle actin, yeast-muscle hybrid actins were constructed (10). Hybrid sub1 actin has muscle-specific residues introduced into the small domain of actin, making it muscle-like in subdomain 1 and yeastlike in subdomains 2, 3, and 4. Sub12 hybrid actin has muscle-specific residues in both subdomains 1 and 2, making it muscle-like in subdomains 1 and 2 and yeastlike in subdomains 3 and 4. These actins showed muscle-like behavior in several biochemical properties: nucleotide exchange rates, thermostability, and the nucleation phase during polymerization. Although most of the residues proposed to be involved in the interaction of actin with myosin are located in subdomains 1 and 2, both hybrids exhibited yeast actin behavior in the activation of myosin ATPase activity. The hybrid actins also exhibited higher critical concentrations than either yeast or muscle actin and formed ADP-F-actin that was less stable than either yeast or muscle actin (10).The experimental results with the hybrid actins indicated that the behavior of actin as a whole is dictated by the interaction and cross-talk of the two halves. One experimental approach to further understand the structural bases of 1) biochemical differences observed with yeast and muscle actins, 2) how introduction of muscle-specific residues affects conformational behavior of the hybrids, and 3) whether there is propagation of change and cross-talk from muscle domains into yeast domains of the hybrid actins in solution is to use NMR. However, NMR as a technique is not suitable here due to the high actin concentrations that would be needed and the propensity of actin to aggregate at these concentrations. Therefore, we utilized hydrogen-deuterium (HD)2 exchange detected with mass spectrometry. This is a powerful technique that can be used to obtain information about conformational changes in proteins in solution on a global as well as a local level.A protein sample placed in a deuterated solution will exchange its amide protons of the polypeptide backbone with exchange times varying from seconds to months (11). The amount and rate of deuterium exchange will depend on factors like temperature, pD, accessible surface area, local environment, and conformational flexibility (1215). Labeled protein samples are digested by pepsin, the protease of choice in HD experiments, because of the requirement for acidic pH in order to minimize back exchange. The resulting peptide fragments are separated by high pressure liquid chromatography and analyzed for deuterium uptake.For a peptide that is undergoing amide proton exchange, the average mass of the peptide will increase with labeling time with an increase in mass greater for the peptides that are more exposed to the solvent. The analysis of deuterium uptake for specific peptides and areas of proteins could potentially provide information about structural and conformational differences between different actin forms. Since changes in conformation among different actin states will affect deuterium uptake, HD exchange coupled with mass spectrometry should provide us with a tool to study the states of actin present in solution. HD exchange detected with mass spectrometry was previously used successfully by Chik and co-workers (16) to assess changes in muscle actin structure due to the G- to F-actin transition and the binding of phallodin to F-actin and of DNase I to G-actin. More recently, the technique has been used to assess conformational changes in the Arp2/3 complex, which contains actin like subunits (17). In this study, we used HD exchange detected with mass spectrometry to analyze differences in solution structures of yeast, muscle, sub1, and sub12 G- and F-actins.  相似文献   

3.
A major factor in profilin regulation of actin cytoskeletal dynamics is its facilitation of G-actin nucleotide exchange. However, the mechanism of this facilitation is unknown. We studied the interaction of yeast (YPF) and human profilin 1 (HPF1) with yeast and mammalian skeletal muscle actins. Homologous pairs (YPF and yeast actin, HPF1 and muscle actin) bound more tightly to one another than heterologous pairs. However, with saturating profilin, HPF1 caused a faster etheno-ATP exchange with both yeast and muscle actins than did YPF. Based on the -fold change in ATP exchange rate/K(d), however, the homologous pairs are more efficient than the heterologous pairs. Thus, strength of binding of profilin to actin and nucleotide exchange rate are not tightly coupled. Actin/HPF interactions were entropically driven, whereas YPF interactions were enthalpically driven. Hybrid yeast actins containing subdomain 1 (sub1) or subdomain 1 and 2 (sub12) muscle actin residues bound more weakly to YPF than did yeast actin (K(d) = 2 microm versus 0.6 microm). These hybrids bound even more weakly to HPF than did yeast actin (K(d) = 5 microm versus 3.2 microm). sub1/YPF interactions were entropically driven, whereas the sub12/YPF binding was enthalpically driven. Compared with WT yeast actin, YPF binding to sub1 occurred with a 5 times faster k(off) and a 2 times faster k(on). sub12 bound with a 3 times faster k(off) and a 1.5 times slower k(on). Profilin controls the energetics of its interaction with nonhybrid actin, but interactions between actin subdomains 1 and 2 affect the topography of the profilin binding site.  相似文献   

4.
We have examined the role of the acidic residues Asp2 and Glu4 at the NH2 terminus of Saccharomyces cerevisiae actin through site-directed mutagenesis. In DNEQ actin, these residues have been changed to Asn2 and Gln4, whereas in delta DSE actin, the Asp2-Ser-Glu tripeptide has been deleted. Both mutant actins can replace wild type yeast actin. Peptide mapping studies reveal that DNEQ, like wild type actin, retains the initiator Met and is NH2 terminally acetylated, whereas delta DSE has a free NH2 terminus and has lost the initiator Met. Interestingly, microscopic examination of filaments of these two actins reveal the appearance of bundled filaments. The DNEQ bundles are smaller and more ordered, whereas the delta DSE bundles are larger and more loosely organized. Additionally, both mutant actins activate the ATPase activity of rabbit muscle myosin S1 fragment to a lesser extent than wild type. We have also developed a sensitive assay for actin function in vivo that enabled us to detect a slight defect in the ability of these mutant actins to support secretion, an important function in yeast. Thus, although the mutant actins resulted in no gross phenotypic changes, we were able to detect a defect in actin function through this assay. From these studies we can conclude that 1) although NH2-terminal negative charges are not essential to yeast life, the loss of such charges does result in a slight defect in the actins' ability to support secretion, 2) removal of the NH2-terminal negative charges promotes the bundling of actin filaments, and 3) actins lacking NH2-terminal negative charges are unable to activate the myosin S1 ATPase activity as well as wild type actin.  相似文献   

5.
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   

6.
In order to investigate the structural basis for functional differences among actin isoforms, we have compared the polymerization properties and conformations of scallop adductor muscle beta-like actin and rabbit skeletal muscle alpha-actin. Polymerization of scallop Ca(2+)-actin was slower than that of skeletal muscle Ca(2+)-actin. Cleavage of the actin polypeptide chain between Gly-42 and Val-43 with Escherichia coli protease ECP 32 impaired the polymerization of scallop Mg(2+)-actin to a greater extent than skeletal muscle Mg(2+)-actin. When monomeric scallop and skeletal muscle Ca(2+)-actins were subjected to limited proteolysis with trypsin, subtilisin, or ECP 32, no differences in the conformation of actin subdomain 2 were detected. At the same time, local differences in the conformations of scallop and skeletal muscle actin subdomains 1 were revealed as intrinsic fluorescence differences. Replacement of tightly bound Ca(2+) with Mg(2+) resulted in more extensive proteolysis of segment 61-69 of scallop actin than in the case of skeletal muscle actin. Furthermore, segment 61-69 was more accessible to proteolysis with subtilisin in polymerized scallop Ca(2+)-actin than in polymerized skeletal muscle Ca(2+)-actin, indicating that, in the polymeric form, the nucleotide-containing cleft is in a more open conformation in beta-like scallop actin than in skeletal muscle alpha-actin. We suggest that this difference between scallop and skeletal muscle actins is due to a less efficient shift of scallop actin subdomain 2 to the position it has in the polymer. The possible consequences of amino acid substitutions in actin subdomain 1 in the allosteric regulation of the actin cleft, and hence in the different stabilities of polymers formed by different actins, are discussed.  相似文献   

7.
Packer LE  Song B  Raleigh DP  McKnight CJ 《Biochemistry》2011,50(18):3706-3712
Villin-type headpiece domains are ~70 residue motifs that reside at the C-terminus of a variety of actin-associated proteins. Villin headpiece (HP67) is a commonly used model system for both experimental and computational studies of protein folding. HP67 is made up of two subdomains that form a tightly packed interface. The isolated C-terminal subdomain of HP67 (HP35) is one of the smallest autonomously folding proteins known. The N-terminal subdomain requires the presence of the C-terminal subdomain to fold. In the structure of HP67, a conserved salt bridge connects N- and C-terminal subdomains. This buried salt bridge between residues E39 and K70 is unusual in a small protein domain. We used mutational analysis, monitored by CD and NMR, and functional assays to determine the role of this buried salt bridge. First, the two residues in the salt bridge were replaced with strictly hydrophobic amino acids, E39M/K70M. Second, the two residues in the salt bridge were swapped, E39K/K70E. Any change from the wild-type salt bridge residues results in unfolding of the N-terminal subdomain, even when the mutations were made in a stabilized variant of HP67. The C-terminal subdomain remains folded in all mutants and is stabilized by some of the mutations. Using actin sedimentation assays, we find that a folded N-terminal domain is essential for specific actin binding. Therefore, the buried salt bridge is required for the specific folding of the N-terminal domain which confers actin-binding activity to villin-type headpiece domains, even though the residues required for this specific interaction destabilize the C-terminal subdomain.  相似文献   

8.
We employed budding yeast and skeletal muscle actin to examine the contribution of the actin isoform to myosin motor function. While yeast and muscle actin are highly homologous, they exhibit different charge density at their N termini (a proposed myosin-binding interface). Muscle myosin-II actin-activated ATPase activity is significantly higher with muscle versus yeast actin. Whether this reflects inefficiency in the ability of yeast actin to activate myosin is not known. Here we optimized the isolation of two yeast myosins to assess actin function in a homogenous system. Yeast myosin-II (Myo1p) and myosin-V (Myo2p) accommodate the reduced N-terminal charge density of yeast actin, showing greater activity with yeast over muscle actin. Increasing the number of negative charges at the N terminus of yeast actin from two to four (as in muscle) had little effect on yeast myosin activity, while other substitutions of charged residues at the myosin interface of yeast actin reduced activity. Thus, yeast actin functions most effectively with its native myosins, which in part relies on associations mediated by its outer domain. Compared with yeast myosin-II and myosin-V, muscle myosin-II activity was very sensitive to salt. Collectively, our findings suggest differing degrees of reliance on electrostatic interactions during weak actomyosin binding in yeast versus muscle. Our study also highlights the importance of native actin isoforms when considering the function of myosins.  相似文献   

9.
Cofilin is a major cytoskeletal protein that binds to both monomeric actin (G-actin) and polymeric actin (F-actin) and is involved in microfilament dynamics. Although an atomic structure of the G-actin-cofilin complex does not exist, models of the complex have been built using molecular dynamics simulations, structural homology considerations, and synchrotron radiolytic footprinting data. The hydrophobic cleft between actin subdomains 1 and 3 and, alternatively, the cleft between actin subdomains 1 and 2 have been proposed as possible high-affinity cofilin binding sites. In this study, the proposed binding of cofilin to the subdomain 1/subdomain 3 region on G-actin has been probed using site-directed mutagenesis, fluorescence labeling, and chemical cross-linking, with yeast actin mutants containing single reactive cysteines in the actin hydrophobic cleft and with cofilin mutants carrying reactive cysteines in the regions predicted to bind to G-actin. Mass spectrometry analysis of the cross-linked complex revealed that cysteine 345 in subdomain 1 of mutant G-actin was cross-linked to native cysteine 62 on cofilin. A cofilin mutant that carried a cysteine substitution in the α3-helix (residue 95) formed a cross-link with residue 144 in actin subdomain 3. Distance constraints imposed by these cross-links provide experimental evidence for cofilin binding between actin subdomains 1 and 3 and fit a corresponding docking-based structure of the complex. The cross-linking of the N-terminal region of recombinant yeast cofilin to actin residues 346 and 374 with dithio-bis-maleimidoethane (12.4 Å) and via disulfide bond formation was also documented. This set of cross-linking data confirms the important role of the N-terminal segment of cofilin in interactions with G-actin.  相似文献   

10.
The Arp2/3 complex creates filament branches leading to an enhancement in the rate of actin polymerization. Work with Arp complexes from different sources indicated that it was inactive by itself, required an activating factor such as the Wiskott-Aldrich syndrome protein (WASP), and might exhibit a preference for ATP or ADP-P(i) actin. However, with yeast actin, P(i) release is almost concurrent with polymerization, eliminating the presence of an ADP-P(i) cap. We thus investigated the ability of the yeast Arp2/3 complex (yArp2/3) to facilitate yeast actin polymerization in the presence and absence of the Arp2/3-activating factor Las17p WA. yArp2/3 significantly accelerates yeast actin but not muscle actin polymerization in the absence of Las17p WA. The addition of Las17p WA further enhances yeast actin polymerization by yArp2/3 and allows the complex to now assist muscle actin polymerization. This actin isoform difference is not observed with bovine Arp2/3 complex, because the neural WASP VCA fragment is required for polymerization of both actins. Observation of individual branching filaments showed that Las17p WA increased the persistence of filament branches. Compared with wild type actin, the V159N mutant actin, proposed to be more ATP-like in behavior, exhibited an enhanced rate of polymerization in the presence of the yArp2/3 complex. yArp2/3 caused a significant rate of P(i) release prior to observation of an increase in filament mass but while branched structures were present. Thus, yeast F-actin can serve as a primary yArp2/3-activating factor, indicating that a newly formed yeast actin filament has a topology, unlike that of muscle actin, that is recognized specifically by yArp2/3.  相似文献   

11.
Differential interactions of tropomyosin (TM) isoforms with actin can be important for determination of the thin filament functions. A mechanism of tropomyosin binding to actin was studied by comparing interactions of five αTM isoforms with actin modified with m-maleimidobenzoyl-N-hydroxysuccinimide ester (MBS) and with fluorescein-5-isothiocyanate (FITC). MBS attachment sites were revealed with mass spectrometry methods. We found that the predominant actin fraction was cross-linked by MBS within subdomain 3. A smaller fraction of the modified actin was cross-linked within subdomain 2 and between subdomains 2 and 1. Moreover, investigated actins carried single labels in subdomains 1, 2, and 3. Such extensive modification caused a large decrease in actin affinity for skeletal and smooth muscle tropomyosins, nonmuscle TM2, and chimeric TM1b9a. In contrast, binding of nonmuscle isoform TM5a was less affected. Isoform’s affinity for actin modified in subdomain 2 by binding of FITC to Lys61 was intermediate between the affinity for native actin and MBS-modified actin except for TM5a, which bound to FITC–actin with similar affinity as to actin modified with MBS. The analysis of binding curves according to the McGhee–von Hippel model revealed that binding to an isolated site, as well as cooperativity of binding to a contiguous site, was affected by both actin modifications in a TM isoform-specific manner.  相似文献   

12.
The fungal toxin cytochalasin D (CD) interferes with the normal dynamics of the actin cytoskeleton by binding to the barbed end of actin filaments. Despite its widespread use as a tool for studying actin-mediated processes, the exact location and nature of its binding to actin have not been previously determined. Here we describe two crystal structures of an expressed monomeric actin in complex with CD: one obtained by soaking preformed actin crystals with CD, and the other obtained by cocrystallization. The binding site for CD, in the hydrophobic cleft between actin subdomains 1 and 3, is the same in the two structures. Polar and hydrophobic contacts play equally important roles in CD binding, and six hydrogen bonds stabilize the actin-CD complex. Many unrelated actin-binding proteins and marine toxins target this cleft and the hydrophobic pocket at the front end of the cleft (viewing actin with subdomain 2 in the upper right corner). CD differs in that it binds to the back half of the cleft. The ability of CD to induce actin dimer formation and actin-catalyzed ATP hydrolysis may be related to its unique binding site and the necessity to fit its bulky macrocycle into this cleft. Contacts with residues lining this cleft appear to be crucial to capping and/or severing. The cocrystallized actin-CD structure also revealed changes in actin conformation. An ∼ 6° rotation of the smaller actin domain (subdomains 1 and 2) with respect to the larger domain (subdomains 3 and 4) results in small changes in crystal packing that allow the D-loop to adopt an extended loop structure instead of being disordered, as it is in most crystal structures of actin. We speculate that these changes represent a potential conformation that the actin monomer can adopt on the pathway to polymerization or in the filament.  相似文献   

13.
The N-terminus of all actins so far studied is acetylated. Although the pathways of acetylation have been well studied, its functional importance has been unclear. A negative charge cluster in the actin N-terminal region is shown to be important for the function of actomyosin. Acetylation at the N-terminus removes a positive charge and increases the amount of net negative charges in the N-terminal region. This may augment the role of the negative charge cluster. To examine this possibility, actin with a nonacetylated N-terminus (nonacetylated actin) was produced. The nonacetylated actin polymerized and depolymerized normally. In actin-activated heavy meromyosin ATPase assays, the nonacetylated actin showed higher K(app) without significantly changing V(max), compared with those of wild-type actin. This is in contrast to the effect of the N-terminal negative charge cluster, which increases V(max) without changing K(app). These results indicate that the acetylation at the N-terminus of actin strengthens weak actomyosin interaction.  相似文献   

14.
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   

15.
Actin is ubiquitous in eukaryotes, nevertheless its existence has not yet been clearly proven in Tetrahymena. Here we report the cloning and sequencing of an actin gene from the genomic library of Tetrahymena pyriformis using a Dictyostelium actin gene as a probe. The Tetrahymena actin gene has no intron. The predicted actin is composed of 375 amino acids like other actins and its molecular weight is estimated as 41,906. Both T. pyriformis and T. thermophila possess a single species of actin genes which differ in their restriction patterns. Northern hybridization analysis revealed that the actin gene was actively transcribed in vivo. To detect the gene product, we synthesized an N-terminal peptide of the deduced sequence and prepared its antibody. Using an immunoblotting technique, we identified Tetrahymena actin on a two-dimensional gel electrophoretic plate. The actin spot migrated near an added spot of rabbit skeletal muscle actin, but clearly differed from the latter in its isoelectric point and apparent molecular weight. The primary structure of Tetrahymena actin shares about 75% homology equally with those of other representative actins. This value is extremely low as a homology rate between known actins. Tetrahymena actin diverges not only in relatively variable regions of other actins, but also in relatively constant regions. The hydrophilicity levels of two regions (residues 190 to 200 and residues 225 to 235) are also quite different between the Tetrahymena actin and skeletal muscle actin. Thus, we conclude that actin is present in Tetrahymena, but it is one of the most unique actins among the actins known hereto.  相似文献   

16.
Diaphanous-related formins (Drf) are activated by Rho GTP binding proteins and induce polymerization of unbranched actin filaments. They contain three formin homology domains. Evidence as to the effect of formins on actin polymerization were obtained using FH2/FH1 constructs of various length from different Drfs. Here we define the core FH2 domain as a proteolytically stable domain of approximately 338 residues. The monomeric FH2 domains from mDia1 and mDia3 inhibit polymerization of actin and can bind in a 1:1 complex with F-actin at micromolar concentrations. The X-ray structure analysis of the domain shows an elongated, crescent-shaped molecule consisting of three helical subdomains. The most highly conserved regions of the domain span a distance of 75 A and are both required for barbed-end inhibition. A construct containing an additional 72 residue linker has dramatically different properties: It oligomerizes and induces actin polymerization at subnanomolar concentration.  相似文献   

17.
Striated muscle thin filaments adopt different quaternary structures, depending upon calcium binding to troponin and myosin binding to actin. Modification of actin subdomain 2 alters troponin-tropomyosin-mediated regulation, suggesting that this region of actin may contain important protein-protein interaction sites. We used yeast actin mutant D56A/E57A to examine this issue. The mutation increased the affinity of tropomyosin for actin 3-fold. The addition of Ca(2+) to mutant actin filaments containing troponin-tropomyosin produced little increase in the thin filament-myosin S1 MgATPase rate. Despite this, three-dimensional reconstruction of electron microscope images of filaments in the presence of troponin and Ca(2+) showed tropomyosin to be in a position similar to that found for muscle actin filaments, where most of the myosin binding site is exposed. Troponin-tropomyosin bound with comparable affinity to mutant and wild type actin in the absence and presence of calcium, and in the presence of myosin S1, tropomyosin bound very tightly to both types of actin. The mutation decreased actin-myosin S1 affinity 13-fold in the presence of troponin-tropomyosin and 2.6-fold in the absence of the regulatory proteins. The results suggest the importance of negatively charged actin subdomain 2 residues 56 and 57 for myosin binding to actin, for tropomyosin-actin interactions, and for regulatory conformational changes in the actin-troponin-tropomyosin complex.  相似文献   

18.
The kinetics and thermodynamics for the polymerization of purified Acanthamoeba actin were studied and compared to muscle actin. Polymerization was qualitatively similar for the two actins with a rate-limiting nucleation step followed by rapid polymer extension. Polymerization occurred only above a threshold critical concentration which varied with polymerization conditions for each actin. In the presence of 2 mM MgCl2, nucleation of both actins was rapid and their critical concentrations were similarly low and not detectably dependent on temperature. In 0.1 M KCl, the rates of nucleation of both actins were much slower than when Mg2+ was present and were significantly different from each other. Also, under these conditions, the critical concentrations of Acanthamoeba and muscle actin were significantly different and both varied markedly with temperature. These quantitative differences between the two actins could be attributed to differences in both their enthalpies and entropies of polymerization, Acanthamoeba actin having the more positive deltaH and delta S. Co-polymerization of the two actins was also demonstrated. Overall, however, there were no qualitative differences between Acanthamoeba and muscle actin that would suggest a unique role for the monomer-polymer equilibrium of cytoplasmic actin in cell motility.  相似文献   

19.
The N-terminal region in actin has been shown to interact with both myosin and troponin (Tn) during the cross-bridge cycle and in regulation. To study the role of this region in regulation, we used yeast actin mutants with increased and decreased numbers of acidic residues. The mutants included D24A/D25A, with Asp(24) and Asp(25) replaced with alanines; DNEQ, with the substitution of Asp(2) and Glu(4) with their amide analogs; and 4Ac, with Glu(3) and Asp(4) inserted in lieu of Ser(3). In the in vitro motility assay, using reconstituted regulated thin filaments, the sliding speeds of DNEQ, D24A/D25A, and 4Ac were similar at all pCa values. Thus, Ca(2+)-sensitivity of the thin filaments and the inhibitory function of TnI appear to be insensitive to changes in charge (+/-2) at the N-terminus of actin, suggesting little, if any, role of that actin region in regulation. A Ca(2+)-independent conformational change in that region was detected upon troponin binding to actin-Tm via an increase in the fluorescence of a pyrene probe attached to another yeast actin mutant that we used (Cys(1)).  相似文献   

20.
Mutations in human alpha-skeletal actin have been implicated in causing congenital nemaline myopathy, a disease characterized histopathologically by nemaline bodies in skeletal muscle and manifested in the patient as skeletal muscle weakness. Here we investigate the functional effects of three severe nemaline myopathy mutations (V43F, A138P, and R183G) in human alpha-skeletal actin. Wild-type and mutant actins were expressed and purified from the baculovirus/insect cell expression system. The mutations are located in different subdomains of actin; Val-43 is located in a flexible loop of subdomain 2, Ala-138 is near a hydrophobic cleft in the "hinge" region between subdomains 1 and 3, and Arg-183 is near the nucleotide-binding site. None of the three mutations affected the folding of the actin monomer, the velocity at which skeletal myosin moves actin in an in vitro motility assay, or the relative average isometric force supported by F-actin. Defects in fundamental actomyosin interactions are, therefore, unlikely to account for the muscle weakness observed in affected patients. There were, however, significant changes observed in the polymerization kinetics of V43F and A138P and in the rate of nucleotide release for V43F. No detectable defect was found for R183G. If these subtle changes in polymerization observed in vitro are amplified in the context of the sarcomere, it could in principle be one of the primary insults that triggers the development of nemaline myopathy.  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