首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Crystal structure studies have shown that cleaved and intact serpins differ essentially in the topology of beta-sheet A. This is five-stranded in the intact molecules and six-stranded after cleavage by insertion of strand s4A whose C-terminus has become free [L?bermann, H., Tokuoka, R., Deisenhofer, J. & Huber, R. (1984) J. Mol. Biol. 177, 531-556; Wright, T. H., Qian, H. X. & Huber, R. (1990) J. Mol. Biol. 213, 513-528]. The structural transition is accompanied by changes in spectral properties and an increase in thermal stability. We show here that an N alpha-acetyl-tetradecapeptide with the amino acid sequence of strand s4A, residues 345-358 of human alpha 1-antitrypsin, associates with intact alpha 1-antitrypsin and forms a stoichiometric complex with properties very similar to cleaved alpha 1-antitrypsin. Complex generation has the characteristics of a folding process.  相似文献   

2.
Purification and characterization of proteins that bind to yeast ARSs   总被引:27,自引:0,他引:27  
Two proteins that bind to yeast ARS DNA have been purified using conventional and oligonucleotide affinity chromatography. One protein has been purified to homogeneity and has a mass of 135 kDa. Competitive binding studies and DNase I footprinting show that the protein binds to a sequence about 80 base pairs away from the core consensus in the region known as domain B. This region has previously been shown to be required for efficient replication of plasmids carrying ARS1 elements. To investigate further whether the protein might have a function related to the ability of ARSs to act as replicators, binding to another ARS was tested. The protein binds to the functional ARS adjacent to the silent mating type locus HMR, called the HMR-E ARS, about 60 base pairs from the core consensus sequence. Surprisingly, there is little homology between the binding site at the HMR-E ARS and the binding site at ARS1. The 135-kDa protein is probably the same as ABF-I (SBF I) (Shore, D., Stillman, D. J. Brand, A. H., and Nasmyth, K. A. (1987) EMBO J. 6, 461-467; Buchman, A. R., Kimmerly, W. J., Rine, J., and Kornberg, R. D. (1988) Mol. Cell. Biol. 8, 210-225). A second DNA-binding protein was separated from ABF-I during later stages of the purification. This protein, which we designate ABF-III, also binds specifically to the ARS1 sequence, as shown by DNase I footprinting, at a site adjacent to the ABF-I recognition site. Purification of these two ARS binding proteins should aid in our understanding of the complex mechanisms that regulate eukaryotic DNA replication.  相似文献   

3.
Crystallographic studies of ribulose-1,5-bisphosphate carboxylase/oxygenase from Rhodospirillum rubrum suggest that active-site Asn111 interacts with Mg2+ and/or substrate (Lundqvist, T., and Schneider, G. (1991) J. Biol. Chem. 266, 12604-12611). To examine possible catalytic roles of Asn111, we have used site-directed mutagenesis to replace it with a glutaminyl, aspartyl, seryl, or lysyl residue. Although the mutant proteins are devoid of detectable carboxylase activity, their ability to form a quaternary complex comprised of CO2, Mg2+, and a reaction-intermediate analogue is indicative of competence in activation chemistry and substrate binding. The mutant proteins retain enolization activity, as measured by exchange of the C3 proton of ribulose bisphosphate with solvent, thereby demonstrating a preferential role of Asn111 in some later step of overall catalysis. The active sites of this homodimeric enzyme are formed by interactive domains from adjacent subunits (Larimer, F. W., Lee, E. H., Mural, R. J., Soper, T. S., and Hartman, F. C. (1987) J. Biol. Chem. 262, 15327-15329). Crystallography assigns Asn111 to the amino-terminal domain of the active site (Knight, S., Anderson, I., and Br?ndén, C.-I. (1990) J. Mol. Biol. 215, 113-160). The observed formation of enzymatically active heterodimers by the in vivo hybridization of an inactive position-111 mutant with inactive carboxyl-terminal domain mutants is consistent with this assignment.  相似文献   

4.
Molecular anatomy of the antibody binding site   总被引:6,自引:0,他引:6  
The binding region of immunoglobulins, which includes the portion of the molecule having the most variability in its amino acid sequence, is shown to have a surprisingly constant structure that can be characterized in terms of a simple, well-defined model. The binding region is composed of the antigen combining site plus its immediate vicinity and arises by noncovalent association of the light and heavy chain variable domains (VL and VH, respectively). The antigen combining site itself consists of six polypeptide chain segments ("hypervariable loops") which comprise some 80 amino acid residues and are attached to a framework of VL and VH beta-sheet bilayers. Having analyzed refined x-ray crystallographic coordinates for three antigen-binding fragments (Fab KOL (Marquart, M., Deisenhofer, J., and Huber, R. (1980) J. Mol. Biol. 141, 369-391), MCPC 603 (Segal, D., Padlan, E. A., Cohen, G. H., Rudikoff, S., Potter, M., and Davies, D. R. (1974) Proc. Natl. Acad. Sci. U. S. A. 71, 4298-4302), and NEW (Saul, F. A., Amzel, L. M., and Poljak, R. J. (1978) J. Biol. Chem. 253, 585-597] we use the results to introduce a general model for the VL-VH interface forming the binding region. The region consists of two closely packed beta-sheets, and its geometry corresponds to a 9-stranded, cylindrical barrel of average radius 0.84 nm with an average angle of -53 degrees between its two constituent beta-sheets. The barrel forms the bottom and sides of the antigen combining site. The model demonstrates that the structural variability of the binding region is considerably less than was thought previously. Amino acid residues which are part of the domain-domain interface and appear not to be accessible to solvent or antigen contribute to antibody specificity.  相似文献   

5.
The binding of Ca2+ to troponin C (TnC) regulates skeletal muscle contraction. We have isolated a full-length cDNA clone for fast skeletal muscle TnC from a neonatal rabbit skeletal muscle library and determined its nucleic acid sequence. The amino acid sequence deduced from this clone matches the previously reported amino acid sequence (Collins, J. H., Greaser, M. L., Potter, J. D., and Horn, M. J. (1977) J. Biol. Chem. 252, 6356-6362) except at the amino terminus. According to the nucleotide sequence, the first 2 residues of TnC are threonine-aspartic acid, which is the reverse of the order reported previously. The isolation of the adult form of TnC from a neonatal library suggests that there may be no developmental isoforms of fast TnC. The protein coding region of the fast TnC clone has 67% homology with the reported nucleotide sequence for chicken slow TnC (Putkey, J. A., Carroll, S. L., and Means, A. R. (1987) Mol. Cell. Biol. 7, 549-1553). The homologies between the nucleotide sequences of TnC, calmodulin, and parvalbumin provide evidence that all three proteins were derived from a common precursor molecule which had four Ca2+-binding sites.  相似文献   

6.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

7.
Previous spectroscopic studies on the phycocyanobilin-containing peptide beta-2T from Synechococcus sp. 6301 C-phycocyanin and the phycoerythrobilin-containing peptide beta-2TP from Porphyridium cruentum B-phycoerythrin indicated a different single thioether mode of attachment, postulated to be through the D-ring of the tetrapyrrole, in contrast to the A-ring linkage established for the other singly linked bilins in these proteins (Bishop, J.E., Lagarias, J.C., Nagy, J. O., Schoenleber, R.W., Rapoport, H., Klotz, A.V., and Glazer, A.N. (1986) J. Biol. Chem. 261, 6790-6796; Klotz, A.V., Glazer, A.N., Bishop, J.E., Nagy, J.O., and Rapoport, H. (1986) J. Biol. Chem. 261, 6797-6805). The crystal structure of Agmenellum quadruplicatum C-phycocyanin at 2.5-A resolution (Schirmer, T., Bode, W., and Huber, R. (1987) J. Mol. Biol., 196, 677-695) supports an A-ring linkage for all three phycocyanobilins. Consequently we have re-evaluated our proposed structural assignments by further 1H NMR studies. Two-dimensional homonuclear correlated and nuclear Overhauser enhancement spectroscopic data presented here show that all three bilins in Synechococcus 6301 C-phycocyanin are attached solely through the A-ring, complementary to the crystallographic data. The evidence from the NMR data for all bilin peptides examined includes the dipoledipole interactions of the 5-H with the 3-H, 3'-H, and a pyrrole methyl group (7-CH3); the corresponding interactions would not be possible in a D-ring-linked bilin. The 5-H also consistently exhibits allylic J-coupling to the 3-H, supporting A-ring linkage assignment. These data are inconsistent with the alternative D-ring linkage assignment since this would involve J-coupling through five bonds. Examination of the phycoerythrobilin beta-2 position in B-phycoerythrin also reveals an A-ring type of attachment by similar criteria. We conclude that all singly linked bilins are attached through the A-ring.  相似文献   

8.
The heme molecular structure of the met-azido form of the myoglobin from the shark Galeorhinus japonicus has been investigated by 1H NMR. A nuclear Overhauser effect (NOE) was clearly observed among the heme peripheral side-chain proton signals of this complex, which undergoes thermal spin equilibrium between high-spin (S = 5/2) and low-spin (S = 1/2) states, and the NOE connectivities provided the assignment of the resonances from the heme C13(1)H2 and C17(1)H2 protons. Chemical shift inequivalence of these proton resonances not only provided information about the orientation of these methylene protons with respect to the heme plane, but also allowed characterization of the time-dependent build-up of the NOE between them, which yields the correlation time for the internal motion of the inter-proton vector. The relatively large mobility found for the C17(1)H2 group suggests that the carboxyl oxygen of the heme C17 propionate is not anchored to the apo-protein by a salt bridge. It has been shown that the ferric high-spin form of G. japonicus Mb possesses a penta-coordinated heme [Suzuki, T. (1987) Biochim. Biophys. Acta 914, 170-176; Yamamoto, Y., Osawa, A., Inoue, Y., Ch?j?, R. & Suzuki, T. (1990) Eur. J. Biochem. 192, 225-229] and that the conformation of both heme propionate groups is fixed with respect to the heme, as well as the apo-protein, by a salt bridge [Yamamoto, Y., Inoue, Y., Ch?j?, R. & Suzuki, T. (1990) Eur. J. Biochem. 189, 567-573]. Therefore the weakening or interruption of the interaction between the C17 propionate and His FG3 upon the changes of the coordination and spin state of the heme iron, during azide ion binding to ferric high-spin G. japonicus Mb, is attributed to the displacement of the FG corner of the apoprotein away from the heme C17 propionate group. A similar structural alteration has been revealed by X-ray structural analyses of unliganded and liganded forms of ferrous hemoproteins [Baldwin, J. & Chothia, C. (1979) J. Mol. Biol. 129, 175-220; Phillips, S.E.V. (1980) J. Mol. Biol. 142, 531-554].  相似文献   

9.
10.
The equilibrium and kinetic properties for the urea-induced unfolding of the alpha subunit of tryptophan synthase from Escherichia coli, Salmonella typhimurium, and five interspecies hybrids were compared to determine the role of protein folding in evolution. The parent proteins differ at 40 positions in the sequence of 268 amino acids, and the hybrids differ by up to 15 amino acids from the Escherichia coli alpha subunit. The results show that all the proteins follow the same folding mechanism and are consistent with a previously proposed hypothesis [Hollecker, M., & Creighton, T. E. (1983) J. Mol. Biol. 168, 409; Krebs, H., Schmid, F. X., & Jaenicke, R. (1983) J. Mol. Biol. 169, 619] that the folding mechanisms are conserved in homologous proteins. Analysis of the kinetic data suggests that the 15 positions at which the parent proteins differ in the amino folding unit, residues 1-188, do not play a role in a rate-limiting step in folding that has been previously identified as the association of the amino and carboxyl folding units [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T. S., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965]. One or more of the 25 positions at which the parent proteins differ in the carboxyl folding unit, residues 189-268, do appear to play a role in this same rate-limiting step.  相似文献   

11.
Equilibrium dialysis measurements of adenine nucleotide binding to chloroplast coupling factor 1 suggest that the enzyme has six binding sites for ADP, adenylyl-beta,gamma-imidodiphosphate (AMP-PNP), and 2'(3')-O-2,4,6-trinitrophenyl-ATP (TNP-ATP). High affinity binding at all six sites requires the divalent cation, Mg2+. Three of the nucleotide-binding sites, sites 1, 2, and 3, have been mapped by fluorescence resonance energy transfer distance measurements (see McCarty, R. E., and Hammes, G. G. (1987) Trends Biochem. Sci. 12, 234-237). Using the same technique, we mapped the location of nucleotide-binding site 4, a tight, exchangeable site (Shapiro, A. B., Huber, A. H., and McCarty, R. E. (1991) J. Biol. Chem. 266, 4194-4200). Two arrangements of the energy transfer map of coupling factor 1 were found which are compatible with the available data. The two arrangements make different predictions about which sites interact in cooperative catalysis.  相似文献   

12.
On the basis of the spatial structure of ascorbate oxidase [Messerschmidt, A., Rossi, A., Ladenstein, R., Huber, R., Bolognesi, M., Gatti, G., Marchesini, A., Petruzzelli, R. & Finazzi-Agro, A. (1989) J. Mol. Biol. 206, 513-529], an alignment of the amino acid sequence of the related blue oxidases, laccase and ceruloplasmin is proposed. This strongly suggests a three-domain structure for laccase closely related to ascorbate oxidase and a six-domain structure of ceruloplasmin. These domains demonstrate homology with the small blue copper proteins. The relationships suggest that laccase, like ascorbate oxidase, has a mononuclear blue copper in domain 3 and a trinuclear copper between domain 1 and 3 and ceruloplasmin has mononuclear copper ions in domains 2, 4 and 6 and a trinuclear copper between domains 1 and 6.  相似文献   

13.
Cartilage-inducing factors A and B (CIF-A and CIF-B) from bovine bone have recently been identified as transforming growth factor-beta (TGF-beta) (Seyedin, S.M., Thompson, A. Y., Bentz, H., Rosen, D. M., McPherson, J. M., Conti, A., Siegel, N. R., Galluppi, G. R., and Piez, K. A. (1986) J. Biol. Chem., 261, 5693-5695) and a unique protein homologous to TGF-beta (Seyedin S. M., Segarini, P. R., Rosen, D. M., Thompson, A. Y., Bentz, H., and Graycar, J. (1987) J. Biol. Chem., 262, 1946-1949), respectively. Although the biological activities of TGF-beta and CIF-B are similar, the divergence of CIF-B from the highly conserved amino acid sequence of TGF-beta prompted an investigation of its receptor binding properties. Three classes of cell surface binding components were identified. Class A has exclusive affinity for TGF-beta; class B has greater affinity for CIF-B; and class C has equal affinity for both proteins. A high molecular weight component, the predominant binding species, was further characterized and shown to consist of two components that are either class B or class C. The differential binding properties of TGF-beta and CIF-B to cell surface components suggest that there are biological activities unique to each of the proteins.  相似文献   

14.
15.
In an extension of earlier studies on the Escherichia coli plasmid-encoded single-stranded DNA-binding proteins pIP71a SSB, F SSB and R64 SSB [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) Biochemistry 26, 3347-3354; Casas-Finet, J. R., Khamis, M. I., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) J. Biol. Chem. 262, 8574-8593], we have investigated the binding of pIP231a SSB to natural and heavy-atom-derivatized single-stranded homopolynucleotides. Fluorimetric equilibrium binding isotherms indicate that pIP231a SSB has a greater solubility at low ionic strength than any other plasmid SSB protein investigated. Furthermore, its complex with mercurated poly(uridylic acid) [poly(Hg5U)] shows a greater resistance to disruption by salt than the other plasmid SSB complexes. Essentially complete binding of pIP231a SSB to poly(Hg5U) could be achieved, and time-resolved optically detected triplet-state magnetic resonance (ODMR) techniques could be applied to the complex. These methods allowed complete resolution of the three Trp chromophores of pIP231a SSB. Comparison of wavelength-selected ODMR results with those obtained for the poly(Hg5U) complex of a point-mutated chromosomal ssb gene product (Eco SSB) carrying substitutions of Phe for Trp [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Murphy, J. B. & Chase, J. W. (1987) J. Biol. Chem. 262, 10938-10945] confirm that Trp40 and Trp54 of pIP231a SSB are stacked in the complex, while Trp88 is not. This is the same distribution of stacked Trp residues found in Eco SSB. These results are confirmed further by specific effects observed on the ODMR signals of pIP231a SSB upon binding to poly(Br5U) and poly(dT), which are known to be caused by the stacking of Trp54 with nucleic acid bases.  相似文献   

16.
The role of solvation on the sequence dependent conformational variabilities in DNA has been studied by calculating hydration free energies from solvent accessible surface areas for several base steps, as a function of various helical parameters, roll, twist and propeller twist. The results of roll calculations suggest opposite trends for AA and GG steps, with the former tending to have a compressed minor groove and the latter a compressed major groove. These trends are consistent with the experimental findings on sequence preferences and the nature of anisotropic bending of DNA observed in nucleosomes (Drew, H.R. and Travers, A.A., J. Mol. Biol. 186, 773-790 (1985); Satchwell, S.C., Drew, H.R. and Travers, A.A., J. Mol. Biol. 191, 659-675 (1986)) and CAP-DNA interactions (Gartenberg, M.R. and Crothers, D.M., Nature 333, 824-829, (1988)). Solvation energy profiles also indicate preferences for the base pairs in GG and AA steps to adopt low and high propeller twists, respectively. Such agreements may either reflect a coincidence of solvation effects with other energy terms or a dominance of solvent effects. The results are discussed in the context of the crystallographic observations of structural tendencies.  相似文献   

17.
Lac repressor (LacR) is a helix-turn-helix motif sequence-specific DNA binding protein. Based on proton NMR spectroscopic investigations, Kaptein and co-workers have proposed that the helix-turn-helix motif of LacR binds to DNA in an orientation opposite to that of the helix-turn-helix motifs of lambda repressor, lambda cro, 434 repressor, 434 cro, and CAP [Boelens, R., Scheek, R., van Boom, J. and Kaptein, R., J. Mol. Biol. 193, 1987, 213-216]. In the present work, we have determined the orientation of the helix-turn-helix motif of LacR in the LacR-DNA complex by the affinity cleaving method. The DNA cleaving moiety EDTA.Fe was attached to the N-terminus of a 56-residue synthetic protein corresponding to the DNA binding domain of LacR. We have formed the complex between the modified protein and the left DNA half site for LacR. The locations of the resulting DNA cleavage positions relative to the left DNA half site provide strong support for the proposal of Kaptein and co-workers.  相似文献   

18.
Limited chymotryptic cleavage of turkey gizzard calponin yields a 13 kDa fragment which could be purified by its ability to bind to Sepharose-immobilized tropomyosin. This 13 kD polypeptide is shown to be derived from a 22 kDa fragment. Complete amino acid sequence analysis of the 13 kD and 22 kD fragments reveals high homology with the formerly characterized smooth muscle-specific protein SM22 alpha (Pearlstone, J.R., Weber, M., Lees-Miller, J.P., Carpenter, M.R. and Smillie L.B., 1987, J. Biol. Chem. 262, 5985-5991) and the product of gene mp20 of Drosophila (Ayme-Southqate, A., Lasko, P., French, C, and Pardue, M.L. [(1989) J. Cell Biol. 108, 521-531]. Futhermore we recognize sequence elements of a putative actin-binding domain of alpha-actinin, the calpactin I or p 36 sequence, and a consensus motif present in the repeats of the gene product of the candidate unc-87 gene of C. elegans (S.D. Goetinck and R.H. Waterston, personal communication).  相似文献   

19.
We have examined the relationship of the ubiquitous 68-70-kDa cytoskeletal-associated protein beta-internexin (Napolitano, E. W., Pachter, J. S., Chin, S. S. M., and Liem, R. K. H. (1985) J. Cell Biol. 101, 1323-1331) to heat-shock cognate 70 (hsc70), the major constitutive member of the mammalian heat-shock protein 70 (hsp70) family of stress proteins. We purify beta-internexin from rat brain microtubules and confirm its identity with hsc70 and the clathrin-uncoating ATPase by the following criteria: 1) The partial sequence of a cyanogen bromide-derived peptide from beta-internexin matches the inferred amino acid sequence of the cDNA clone pRC62 encoding hsc70 from rat brain (O'Malley, K., Mauron, A., Barchas, J. D., and Kedes, L. (1985) Mol. Cell. Biol. 5, 3476-3483). 2) Mixing experiments followed by two-dimensional gel analyses reveal the precise co-migration of beta-internexin, the clathrin-uncoating ATPase, and the in vitro translation product of cDNA clone pHSP-4 encoding rat brain hsc70. 3) beta-Internexin is recognized by a monoclonal antibody reactive against the class of hsp70 proteins. 4) beta-Internexin purified from a microtubule-associated protein-enriched fraction of rat brain by virtue of high affinity binding to ATP-agarose possesses clathrin cage-specific ATPase activity.  相似文献   

20.
The regulation of striated muscle contraction involves changes in the interactions of troponin and tropomyosin with actin thin filaments. In resting muscle, myosin-binding sites on actin are thought to be blocked by the coiled-coil protein tropomyosin. During muscle activation, Ca2+ binding to troponin alters the tropomyosin position on actin, resulting in cyclic actin-myosin interactions that accompany muscle contraction. Evidence for this steric regulation by troponin-tropomyosin comes from X-ray data [Haselgrove, J.C., 1972. X-ray evidence for a conformational change in the actin-containing filaments of verterbrate striated muscle. Cold Spring Habor Symp. Quant. Biol. 37, 341-352; Huxley, H.E., 1972. Structural changes in actin and myosin-containing filaments during contraction. Cold Spring Habor Symp. Quant. Biol. 37, 361-376; Parry, D.A., Squire, J.M., 1973. Structural role of tropomyosin in muscle regulation: analysis of the X-ray diffraction patterns from relaxed and contracting muscles. J. Mol. Biol. 75, 33-55] and electron microscope (EM) data [Spudich, J.A., Huxley, H.E., Finch, J., 1972. Regulation of skeletal muscle contraction. II. Structural studies of the interaction of the tropomyosin-troponin complex with actin. J. Mol. Biol. 72, 619-632; O'Brien, E.J., Gillis, J.M., Couch, J., 1975. Symmetry and molecular arrangement in paracrystals of reconstituted muscle thin filaments. J. Mol. Biol. 99, 461-475; Lehman, W., Craig, R., Vibert, P., 1994. Ca2+-induced tropomyosin movement in Limulus thin filaments revealed by three-dimensional reconstruction. Nature 368, 65-67] each with its own particular strengths and limitations. Here we bring together some of the latest information from EM analysis of single thin filaments from Pirani et al. [Pirani, A., Xu, C., Hatch, V., Craig, R., Tobacman, L.S., Lehman, W. (2005). Single particle analysis of relaxed and activated muscle thin filaments. J. Mol. Biol. 346, 761-772], with synchrotron X-ray data from non-overlapped muscle fibres to refine the models of the striated muscle thin filament. This was done by incorporating current atomic-resolution structures of actin, tropomyosin, troponin and myosin subfragment-1. Fitting these atomic coordinates to EM reconstructions, we present atomic models of the thin filament that are entirely consistent with a steric regulatory mechanism. Furthermore, fitting the atomic models against diffraction data from skinned muscle fibres, stretched to non-overlap to preclude crossbridge binding, produced very similar results, including a large Ca2+-induced shift in tropomyosin azimuthal location but little change in the actin structure or apparent alteration in troponin position.  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