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 39 毫秒
1.
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   

2.
The primary structure of all actins except that isolated from Naegleria gruberi contains a unique N tau-methylhistidine (MeHis) at position 73. This modified residue has been implicated as possibly being important for the post-translational processing of actin's amino terminus, the binding of actin to DNase I, and in the polymerization of G-actin. We have investigated the potential role of MeHis in each of these processes by utilizing site-directed mutagenesis to change His-73 of skeletal muscle actin to Arg and Tyr. Wild type and mutant actins were synthesized in vivo, using non-muscle cells transfected with mutant cDNAs, and in vitro by translating mutant RNAs synthesized using SP6 RNA polymerase in a rabbit reticulocyte lysate. We have found that actins containing Arg or Tyr at position 73 undergo amino-terminal processing, bind to DNase I-agarose, and become incorporated into the cytoskeleton of a nonmuscle cell as efficiently as wild type actin. Furthermore, using an in vitro copolymerization assay we have found that although there is no difference between the Arg mutant and the wild type actins, the Tyr mutant has a slightly greater critical concentration for polymerization. These results show that MeHis is not absolutely required for any of these processes.  相似文献   

3.
All but 11 of the 323 known actin sequences have Tyr at position 53, and the 11 exceptions have the conservative substitution Phe, which raises the following questions. What is the critical role(s) of Tyr-53, and, if it can be replaced by Phe, why has this happened so infrequently? We compared the properties of purified endogenous Dictyostelium actin and mutant constructs with Tyr-53 replaced by Phe, Ala, Glu, Trp, and Leu. The Y53F mutant did not differ significantly from endogenous actin in any of the properties assayed, but the Y53A and Y53E mutants differed substantially; affinity for DNase I was reduced, the rate of nucleotide exchange was increased, the critical concentration for polymerization was increased, filament elongation was inhibited, and polymerized actin was in the form of small oligomers and imperfect filaments. Growth and/or development of cells expressing these actin mutants were also inhibited. The Trp and Leu mutations had lesser but still significant effects on cell phenotype and the biochemical properties of the purified actins. We conclude that either Tyr or Phe is required to maintain the functional conformations of the DNase I-binding loop (D-loop) in both G- and F-actin, and that the conformation of the D-loop affects not only the properties that directly involve the D-loop (binding to DNase I and polymerization) but also allosterically modifies the conformation of the nucleotide-binding cleft, thus increasing the rate of nucleotide exchange. The apparent evolutionary “preference” for Tyr at position 53 may be the result of Tyr allowing dynamic modification of the D-loop conformation by phosphorylation (Baek, K., Liu, X., Ferron, F., Shu, S., Korn, E. D., and Dominguez, R. (2008) Proc. Natl. Acad. Sci. U.S.A. 105, 11748–11753) with effects similar, but not identical, to those of the Ala and Glu mutations.  相似文献   

4.
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   

5.
The actin monomer-binding protein, profilin, influences the dynamics of actin filaments in vitro by suppressing nucleation, enhancing nucleotide exchange on actin, and promoting barbed-end assembly. Profilin may also link signaling pathways to actin cytoskeleton organization by binding to the phosphoinositide PIP(2) and to polyproline stretches on several proteins. Although activities of profilin have been studied extensively in vitro, the significance of each of these activities in vivo needs to be tested. To study profilin function, we extensively mutagenized the Saccharomyces cerevisiae profilin gene (PFY1) and examined the consequences of specific point mutations on growth and actin organization. The actin-binding region of profilin was shown to be critical in vivo. act1-157, an actin mutant with an increased intrinsic rate of nucleotide exchange, suppressed defects in actin organization, cell growth, and fluid-phase endocytosis of pfy1-4, a profilin mutant defective in actin binding. In reactions containing actin, profilin, and cofilin, profilin was required for fast rates of actin filament turnover. However, Act1-157p circumvented the requirement for profilin. Based on the results of these studies, we conclude that in living cells profilin promotes rapid actin dynamics by regenerating ATP actin from ADP actin-cofilin generated during filament disassembly.  相似文献   

6.
Mutational analysis of yeast profilin.   总被引:13,自引:2,他引:11       下载免费PDF全文
We have mutated two regions within the yeast profilin gene in an effort to functionally dissect the roles of actin and phosphatidylinositol 4,5-bisphosphate (PIP2) binding in profilin function. A series of truncations was carried out at the C terminus of profilin, a region that has been implicated in actin binding. Removal of the last three amino acids nearly eliminated the ability of profilin to bind polyproline in vitro but had no dramatic in vivo effects. Thus, the extreme C terminus is implicated in polyproline binding, but the physiological relevance of this interaction is called into question. More extensive truncation, of up to eight amino acids, had in vivo effects of increasing severity and resulted in changes in conformation and expression level of the mutant profilins. However, the ability of these mutants to bind actin in vitro was not eliminated, suggesting that this region cannot be solely responsible for actin binding. We also mutagenized a region of profilin that we hypothesized might be involved in PIP2 binding. Alteration of basic amino acids in this region produced mutant profilins that functioned well in vivo. Many of these mutants, however, were unable to suppress the loss of adenylate cyclase-associated protein (Cap/Srv2p [A. Vojtek, B. Haarer, J. Field, J. Gerst, T. D. Pollard, S. S. Brown, and M. Wigler, Cell 66:497-505, 1991]), indicating that a defect could be demonstrated in vivo. In vitro assays demonstrated that the inability to suppress loss of Cap/Srv2p correlated with a defect in the interaction with actin, independently of whether PIP2 binding was reduced. Since our earlier studies of Acanthamoeba profilins suggested the importance of PIP2 binding for suppression, we conclude that both activities are implicated and that an interplay between PIP2 binding and actin binding may be important for profilin function.  相似文献   

7.
HUT-14 cells, tumorigenic human fibroblasts, express a mutant beta-actin which has a single amino acid substitution at position 244 (glycine to aspartic acid), in addition to normal beta- and gamma-actin. In order to characterize the biochemical function of the mutant beta-actin, actins were extracted and purified from HUT-14 cells. The partially purified actin fraction contained beta-, gamma-, and mutant beta-actins in the ratio of 1:1:1, the same ratio as in the cells. When the actin of this fraction was purified through a polymerization step, mutant beta-actin was always less incorporated into actin filaments than beta- and gamma-actin. When the polymerization ability of purified HUT-14 actins was examined by sedimentation technique, it was lower than those of muscle and of cytoplasmic actins from another human cell line (HUT-11) which expresses only normal beta- and gamma-actin, in the ratio of 2:1. The deficient polymerization of mutant beta-actin was also observed by examining the ratio of beta-, gamma-, and mutant beta-actins incorporated into actin filaments. The ratio of mutant beta-actin in polymerized actins under all conditions examined was always less than that before polymerization. These results indicate that the single amino acid substitution at position 244 caused the reduction of incorporation of the mutant beta-actin into actin filaments in vitro.  相似文献   

8.
Most actins examined to date undergo a unique posttranslational modification termed processing, catalyzed by the actin N-acetylaminopeptidase. Processing is the removal of acetylmethionine from the amino terminus in class I actins with Met-Asp(Glu) amino termini. For class II actins with Met-X-Asp(Glu) amino termini, processing is the removal of the second residue as an N-acetylamino acid. Other cytosolic proteins with these amino termini are not processed suggesting that the reaction may be specific for actins. In actin, X is usually cysteine. However, there are some class II actins in which this residue is other than cysteine, suggesting a broader substrate specificity for actin N-acetylaminopeptidase than acetylmethionine or acetylcysteine. We constructed mutant actins in which this cysteine was replaced with serine, asparagine, glycine, aspartic acid, histidine, phenylalanine, and tyrosine and used these to determine the substrate specificity of rat liver actin N-acetylaminopeptidase in vitro. Amino-terminal acetylmethinonine was cleaved from adjacent aspartic acid, asparagine, or histidine, but not serine, glycine, phenylalanine, or tyrosine. Of the acetylated actin amino termini tested, only acetylmethionine and acetylcysteine were cleaved. Histidine was never N-acetylated and was not cleaved. When phenylalanine and tyrosine were adjacent to the initiator methionine, no initiator methionine was cleaved even though it was acetylated. These results suggest a narrow substrate specificity for the rat liver actin N-acetylaminopeptidase. They also demonstrate that the adjacent residue can effect actin N-acetylaminopeptidase specificity.  相似文献   

9.
A macronuclear gene-sized molecule carrying an actin gene from the hypotrich ciliate, Histriculus cavicola, was characterized. Southern blot analysis using a coding region probe suggested that actin in H. cavicola is encoded by a single gene. A comparison of the promoter regions indicated that the H. cavicola actin gene has a TATA box in the 5' flanking region in a position identical to those in other oxytrich ciliates. The coding sequence of this gene is not interrupted by any introns, and codes for a protein of 375 amino acid residues. This protein shares a high degree of similarity with other oxytrichid actins, and a relatively low similarity with actins from other eukaryotes. Comparative analyses of sequences indicated that most of the amino acid substitutions in hypotrich actins are found in surface loops, while the core structures are well-conserved. The sites that interact with DNase I and several regions involved in actin-actin contact have diverged considerably in hypotrich actins, while nucleotide-binding sites are the best-conserved interaction motif.  相似文献   

10.
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   

11.
E Prochniewicz  D D Thomas 《Biochemistry》2001,40(46):13933-13940
We have examined the effects of actin mutations on myosin binding, detected by cosedimentation, and actin structural dynamics, detected by spectroscopic probes. Specific mutations were chosen that have been shown to affect the functional interactions of actin and myosin, two mutations (4Ac and E99A/E100A) in the proposed region of weak binding to myosin and one mutation (I341A) in the proposed region of strong binding. In the absence of nucleotide and salt, S1 bound to both wild-type and mutant actins with high affinity (K(d) < microM), but either ADP or increased ionic strength decreased this affinity. This decrease was more pronounced for actins with mutations that inhibit functional interaction with myosin (E99A/E100A and I341A) than for a mutation that enhances the interaction (4Ac). The mutations E99A/E100A and I341A affected the microsecond time scale dynamics of actin in the absence of myosin, but the 4Ac mutation did not have any effect. The binding of myosin eliminated these effects of mutations on structural dynamics; i.e., the spectroscopic signals from mutant actins bound to S1 were the same as those from wild-type actin. These results indicate that mutations in the myosin binding sites affect structural transitions within actin that control strong myosin binding, without affecting the structural dynamics of the strongly bound actomyosin complex.  相似文献   

12.
We have studied the importance of N-terminal processing for normal actin function using the Drosophila Act88F actin gene transcribed and translated in vitro. Despite having different charges as determined by two-dimensional (2D) gel electrophoresis, Act88F expressed in vivo and in vitro in rabbit reticulocyte lysate bind to DNase I with equal affinity and are able to copolymerise with bulk rabbit actin equally well. Using peptide mapping and thin-layer electrophoresis we have shown that bestatin [( 3-amino-2-hydroxy-4-phenyl-butanoyl]-L-leucine), an inhibitor of aminopeptidases, can inhibit actin N-terminal processing in rabbit reticulocyte lysate. Although processed and unprocessed actins translated in vitro are able to bind to DNase I equally well, unprocessed actins are less able to copolymerise with bulk actins. This effect is more pronounced when bulk rabbit actin is used but is still seen with bulk Lethocerus actin. Also, the unprocessed actins reduce the polymerisation of the processed actin translated in vitro with the bulk rabbit actin. This suggests that individual actins do interact, even in non-polymerising conditions. The reduced ability of unprocessed actin to polymerise shows that correct post-translational modification of the N terminus is required for normal actin function.  相似文献   

13.
S Suetsugu  H Miki    T Takenawa 《The EMBO journal》1998,17(22):6516-6526
Profilin was first identified as an actin monomer binding protein; however, recent reports indicate its involvement in actin polymerization. To date, there is no direct evidence of a functional role in vivo for profilin in actin cytoskeletal reorganization. Here, we prepared a profilin mutant (H119E) defective in actin binding, but retaining the ability to bind to other proteins. This mutant profilin I suppresses actin polymerization in microspike formation induced by N-WASP, the essential factor in microspike formation. Profilin associates both in vivo and in vitro with N-WASP at proline-rich sites different from those to which Ash/Grb2 binds. This association between profilin and N-WASP is required for N-WASP-induced efficient microspike elongation. Moreover, we succeeded in reconstituting microspike formation in permeabilized cells using profilin I combined with N-WASP and its regulator, Cdc42. These findings provide the first evidence that profilin is a key molecule linking a signaling network to rapid actin polymerization in microspike formation.  相似文献   

14.
Two IgG1, kappa monoclonal antibodies (Mab) against actin have been obtained from a fusion in which chicken gizzard actin was used as the immunogen. One Mab, designated B4, shows a preferential reactivity toward enteric smooth muscle actin but also cross-reacts with skeletal, cardiac, and aorta actins on the basis of immunoblots, ELISA assays, and indirect immunofluorescence. However, this antibody does not react with either cytoplasmic actin in any of these assay systems. A second Mab, designated C4, reacts with all six known vertebrate isoactins as well as Dictyostelium discoideum and Physarum polycephalum actins. Thus B4 Mab appears to react with an epitope that is at least partially shared among the muscle actins but not found in cytoplasmic actins, while C4 Mab binds to an antigenic determinant that has been highly conserved among the actins. The binding sites of both Mabs on skeletal actin overlap that of pancreatic DNase I. Both antibodies bind a SV8 proteolytic product comprising the amino-terminal two-thirds of the actin molecule, and their epitopes appear to overlap since C4 can compete for the binding of B4 to skeletal actin. Neither antibody is able to prevent actin polymerization.  相似文献   

15.
Heat-shock proteins (hsps) are constitutively induced by the mutant actins in the Drosophila indirect flight muscles (IFM). We compared primary structures of the mutant actin genes (KM75 and HH5) which induce hsps and of the non-inducing alleles (KM129 and KM88). The KM75 actin has lost 20 amino acids at the C-terminus. The HH5 actin has only one amino acid substitution, from Gly-336 to Ser. In KM129, the C-terminal part of actin is replaced by novel amino acids. KM88 is a null allele, with an amber mutation early in the coding region of the mutated actin gene. Although all of the KM75, HH5 and KM129 actins have defects near the C-terminus, only hsp-inducing mutant actins cause enlargement of the IFM nuclei as well as a disruption of myofibrils even in the presence of two copies of the normal genes. We further consider the underlying mechanisms linking these features of the hsp-inducing alleles.  相似文献   

16.
Actin was extracted from isolated nuclei of Acanthamoeba castellanii and purified to homogeneity under nondenaturing conditions by diethylaminoethylcellulose and Sephadex G-100 chromatography. The pure protein has the same molecular weight as cytoplasmic Acanthamoeba actin and a very similar amino acid composition. Isoelectrofocusing shows that nuclear actin is slightly more acidic than the major cytoplasmic species, and comparative analysis of peptides from tryptic and cyanogen bromide digests shows that both actins are very similar but not chemically identical. In an assay that is specific for most actins, the inhibition of DNase I through the formation of a 1:1 G-actin-DNase I complex, the nuclear and cytoplasmic actins are equally effective. By use of a similar procedure for the purification of both actins, it is estimated that the amount of nuclear actin is about 1.5% of the amount of cytoplasmic actin, a major protein of the amoeba. It is concluded that a minor isoelectric species of actin associates selectively with the nuclei of A. castellanii.  相似文献   

17.
ABSTRACT The full length coding sequence of the Euglena gracilis actin gene was determined by RT-PCR of Euglena gracilis mRNA. Conserved regions in the actin amino acid sequence were used as guides for the synthesis of degenerate primers. Sequence was obtained for 1.238 nucleotides, of which 1.131 were coding for 377 amino acids. Sequence comparisons showed a similarity with other actins of 56% to 80%. Even though most of the actin amino acid sequence was conserved, some regions showed high divergence, i.e. the DNase I-binding loop at the N-terminal region. The construction of a phylogenetic tree based on actin sequences from different organisms placed Euglena gracilis in a cluster with Trypanosoma brucei and Leishmania major.  相似文献   

18.
19.
Vacuolar H(+)-ATPase (V-ATPase) binds microfilaments, and that interaction may be mediated by an actin binding domain in subunit B of the enzyme. To test for possible physiologic functions of the actin binding activity of V-ATPase, early responses of resorbing osteoclasts to inhibition of phosphatidylinositol 3-kinase activity by wortmannin and LY294002 were examined. Rapid co-localization between V-ATPase and F-actin was demonstrated by immunocytochemistry, and corresponding association between V-ATPase and F-actin in immunoprecipitations and pelleting assays was detected. This response was reversed as osteoclasts recovered resorptive activity after inhibitors were removed. By expressing and characterizing fusion proteins containing segments of the actin-binding amino-terminal regions of the B subunits of V-ATPase, we mapped the actin-binding site to a 44-amino acid domain. An 11-amino acid segment with a sequence similar to the actin-binding site of human profilin I was detected within this region. 13-Mers containing these profilin-like segments bound actin in fluorescent anisotropy studies and competed with profilin for binding to actin. Using site-directed mutagenesis, the 11-amino acid profilin-like actin-binding motifs (amino acids 49-59 of B1 and 55-65 of B2) were replaced with an 11-amino acid spacer with a sequence based on the homologous sequence from subunit B of Pyrococcus horikoshii, an organism that lacks an actin cytoskeleton. These substitutions eliminated the actin-binding activity of the B subunit fusion proteins. In summary, binding between V-ATPase and F-actin in osteoclasts occurs in response to blocking phosphatidylinositol 3-kinase activity. This response was fully reversible. The actin binding activities of the B subunits of V-ATPase required 11-amino acid actin-binding motifs that are similar in sequence to the actin-binding site of mammalian profilin I.  相似文献   

20.
Unusual metabolism of the yeast actin amino terminus   总被引:5,自引:0,他引:5  
In this paper we have examined the post-translational modifications of the NH2 terminus of actin from the yeast Saccharomyces cerevisiae. Like actins examined previously, this actin contains an acetylated NH2 terminus. Actins in other organisms undergo a unique post-translational processing event in which the initial amino acid(s) are removed by an actin-specific processing enzyme in an acetylation-dependent reaction. This is defined as actin processing. In yeast, actin retains its initiator Met in vivo and is thus not processed even though a rat liver actin processing enzyme can process yeast actin in vitro. This lack of actin processing appears to be a general property of fungi, as the actin from three other species, Aspergillus nidulans, Schizosaccharomyces pombe, and Candida albicans are not NH2 terminally processed either. Yeast actin is a class I actin; its initiator Met directly precedes an acidic residue. We converted yeast actin to a class II species by inserting a Cys codon between the Met-1 and Asp-2 codons. In normal class II actins the Cys residue is removed as acetyl-Cys during processing. Neither the mutant actin nor chick beta-actin (a class I actin) are processed when expressed in yeast. S. cerevisiae thus appears to be also incapable of processing exogenous actins. Further study of the mutant actin containing a Cys at position 2 shows that 30-40% of this actin is stably unacetylated. This unacetylated actin does not have a shorter half-life than the acetylated form. From these studies we conclude that 1) NH2-terminal actin-specific processing is not required for actin function in yeast and three other fungi, 2) yeast are apparently incapable of processing any type of actin precursor, and 3) the stability of a yeast pseudo-class II actin is not affected by the acetylation state of the NH2 terminus.  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