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 31 毫秒
1.
Fluorescence energy transfer was measured between Cys-10 residues in an F-actin filament using 5-[2-((iodoacetyl)amino)-ethyl]aminonaphthalene-1-sulphonic acid (1,5-IAEDANS) as a fluorescence energy donor and 4-dimethylaminophenylazophenyl-4'-maleimide (DABMI) as the acceptor. Both labels were covalently attached to Cys-10 residues in an F-actin filament. Taking the helical structure of the F-actin filament into consideration, the radial coordinate of Cys-10 was calculated to be 23 A. This corresponds to a distance between adjacent sites along the long pitch helix of 56.1 A and along the genetic helix of 53.3 A.  相似文献   

2.
The nucleotide binding site in actin was occupied with the fluorescent analogue formycin A 5' triphosphate which acted as a fluorescent donor for the acceptor chromophore dansyl chloride attached to Tyr-69. The distance separating the two chromophores was calculated to be 2.1 nm from the fluorescence energy transfer measurements. Similar measurements were made of the distances separating dansyl chloride, acting as donor, on Tyr-69 from Co2+ occupying the metal binding site. A distance of 2.1 nm was similarly obtained.  相似文献   

3.
Fluorescence resonance energy transfer (FRET) was used to construct an atomic model of the actin–tropomyosin (Tm) complex on a reconstituted thin filament. We generated five single-cysteine mutants in the 146–174 region of rabbit skeletal muscle α-Tm. An energy donor probe was attached to a single-cysteine Tm residue, while an energy acceptor probe was located in actin Gln41, actin Cys374, or the actin nucleotide binding site. From these donor–acceptor pairs, FRET efficiencies were determined with and without Ca2+. Using the atomic coordinates for F-actin and Tm, we searched all possible arrangements for Tm segment 146–174 on F-actin to calculate the FRET efficiency for each donor–acceptor pair in each arrangement. By minimizing the squared sum of deviations for the calculated FRET efficiencies from the observed FRET efficiencies, we determined the location of the Tm segment on the F-actin filament. Furthermore, we generated a set of five single-cysteine mutants in each of the four Tm regions 41–69, 83–111, 216–244, and 252–279. Using the same procedures, we determined each segment's location on the F-actin filament. In the best-fit model, Tm runs along actin residues 217–236, which were reported to compose the Tm binding site. Electrostatic, hydrogen-bonding, and hydrophobic interactions are involved in actin and Tm binding. The C-terminal region of Tm was observed to contact actin more closely than did the N-terminal region. Tm contacts more residues on actin without Ca2+ than with it. Ca2+-induced changes on the actin–Tm contact surface strongly affect the F-actin structure, which is important for muscle regulation.  相似文献   

4.
F-Actin was digested with alpha-chymotrypsin in 6 M urea, and two peptide fragments from subdomain 4 of actin molecule [Kabsch, W., Mannherz, H.G., Suck, D., Pai, E.F., & Holmes K.C. (1990) Nature 347, 37-44] were purified by reverse-phase HPLC and Sephadex G-50 gel filtration. The peptide fragments were identified as segments from Arg-177 to Tyr-198 (2.6-kDa peptide) and from Ser-199 to Tyr-279 (9.1-kDa peptide). Their effects on actin polymerization induced by 50 or 100 mM KCl were studied by measuring the increase in viscosity by the falling ball method. The 2.6-kDa peptide decreased the rate of actin polymerization and increased the critical concentration for the polymerization. Based on the atomic model of the actin filament [Holmes, K.C., Popp, D., Gebhard, W., & Kabsch, W. (1990) Nature 347, 44-49], the peptide is presumed to bind to the barbed end of the actin filament and inhibit the polymerization. By assuming that the peptide affected the rate of association of the actin monomer to the end of the actin filament, well-fitting curves for the polymerization kinetics were calculated. Computer-assisted results indicated that the dissociation constant of the 2.6-kDa peptide for F-actin is 200 to 260 microM. In contrast, the 9.1-kDa peptide only slightly inhibited actin polymerization. These results suggest that the actin-actin interface in the region between Arg-177 and Tyr-198 has a stronger interaction than those between Ser-199 and Tyr-279. The amino acid sequence L-T-D-Y-L present in the 2.6-kDa segment is homologous to a common sequence in the F-actin capping domain of various actin-binding proteins.  相似文献   

5.
R67 dihydrofolate reductase (R67 DHFR) catalyzes the transfer of a hydride ion from NADPH to dihydrofolate, generating tetrahydrofolate. The homotetrameric enzyme provides a unique environment for catalysis as both ligands bind within a single active site pore possessing 222 symmetry. Mutation of one active site residue results in concurrent mutation of three additional symmetry-related residues, causing large effects on binding of both ligands as well as catalysis. For example, mutation of symmetry-related tyrosine 69 residues to phenylalanine (Y69F), results in large increases in Km values for both ligands and a 2-fold rise in the kcat value for the reaction (Strader, M. B., Smiley, R. D., Stinnett, L. G., VerBerkmoes, N. C., and Howell, E. E. (2001) Biochemistry 40, 11344-11352). To understand the interactions between specific Tyr-69 residues and each ligand, asymmetric Y69F mutants were generated that contain one to four Y69F mutations. A general trend observed from isothermal titration calorimetry and steady-state kinetic studies of these asymmetric mutants is that increasing the number of Y69F mutations results in an increase in the Kd and Km values. In addition, a comparison of steady-state kinetic values suggests that two Tyr-69 residues in one half of the active site pore are necessary for NADPH to exhibit a wild-type Km value. A tyrosine 69 to leucine mutant was also generated to approach the type(s) of interaction(s) occurring between Tyr-69 residues and the ligands. These studies suggest that the hydroxyl group of Tyr-69 is important for interactions with NADPH, whereas both the hydroxyl group and hydrophobic ring atoms of the Tyr-69 residues are necessary for proper interactions with dihydrofolate.  相似文献   

6.
The actin microfilament (F-actin) is a structural and functional component of the cell cytoskeleton. Notwithstanding the primary role it plays for the mechanics of the cell, the mechanical behaviour of F-actin is still not totally explored. In particular, the relationship between the mechanics of F-actin and its molecular architecture is not completely understood. In this study, the mechanical properties of F-actin were related to the molecular topology of its building monomers (G-actin) by employing a computational multi-level approach. F-actins with lengths up to 500 nm were modelled and characterized, using a combination of equilibrium molecular dynamics (MD) simulations and normal mode analysis (NMA). MD simulations were performed to analyze the molecular rearrangements of G-actin in physiological conditions; NMA was applied to compute the macroscopic properties of F-actin from its vibrational modes of motion. Results from this multi-level approach showed that bending stiffness, bending modulus and persistence length are independent from the length of F-actin. On the contrary, the orientations and motions of selected groups of residues of G-actin play a primary role in determining the filament flexibility. In conclusion, this study (i) demonstrated that a combined computational approach of MD and NMA allows to investigate the biomechanics of F-actin taking into account the molecular topology of the filament (i.e., the molecular conformations of G-actin) and (ii) that this can be done using only crystallographic G-actin, without the need of introducing experimental parameters nor of reducing the number of residues.  相似文献   

7.
L D Burtnick  K Bhangu 《FEBS letters》1986,198(2):307-310
5-[Dimethylamino]naphthalene-1-sulfonyl chloride (dansyl chloride) reacts with rabbit skeletal muscle tropomyosin (TM) to yield a highly fluorescent product, DNS-TM. The extent of modification of TM can be regulated over a wide range, 0.3-15.5 dansyl groups per TM, depending upon the temperature and duration of the reaction. However, under all conditions employed, about 15 different fluorescent tryptic peptides of TM were produced. DNS-TM undergoes end-to-end polymerization at low ionic strengths, but to a somewhat lesser extent than unlabelled TM does. DNS-TM also binds muscle F-actin. This interaction may be followed fluorimetrically by observing a blue-shift in emission maximum, an increase in fluorescence intensity or an increase in fluorescence polarization of the DNS-TM complex with F-actin.  相似文献   

8.
Fluorescence energy transfer between nucleotide binding sites in an F-actin filament was measured using 1-N6-ethenoadenosine diphosphate (epsilon-ADP) as a fluorescent donor and 2'(or 3')-O-(2,4,6-trinitrophenyl)adenosine 5'-diphosphate (TNP-ADP) as an acceptor, both of which were bound to F-actin. Taking into consideration the helical structure of the F-actin filament, the radial coordinate of the nucleotide binding site was calculated to be 25 A, which corresponds to a distance between these sites along the long-pitch helix of 56.3 A and along the genetic helix of 56.7 A.  相似文献   

9.
The conformational dynamics of filamentous actin (F-actin) is essential for the regulation and functions of cellular actin networks. The main contribution to F-actin dynamics and its multiple conformational states arises from the mobility and flexibility of the DNase I binding loop (D-loop; residues 40-50) on subdomain 2. Therefore, we explored the structural constraints on D-loop plasticity at the F-actin interprotomer space by probing its dynamic interactions with the hydrophobic loop (H-loop), the C-terminus, and the W-loop via mutational disulfide cross-linking. To this end, residues of the D-loop were mutated to cysteines on yeast actin with a C374A background. These mutants showed no major changes in their polymerization and nucleotide exchange properties compared to wild-type actin. Copper-catalyzed disulfide cross-linking was investigated in equimolar copolymers of cysteine mutants from the D-loop with either wild-type (C374) actin or mutant S265C/C374A (on the H-loop) or mutant F169C/C374A (on the W-loop). Remarkably, all tested residues of the D-loop could be cross-linked to residues 374, 265, and 169 by disulfide bonds, demonstrating the plasticity of the interprotomer region. However, each cross-link resulted in different effects on the filament structure, as detected by electron microscopy and light-scattering measurements. Disulfide cross-linking in the longitudinal orientation produced mostly no visible changes in filament morphology, whereas the cross-linking of D-loop residues > 45 to the H-loop, in the lateral direction, resulted in filament disruption and the presence of amorphous aggregates on electron microscopy images. A similar aggregation was also observed upon cross-linking the residues of the D-loop (> 41) to residue 169. The effects of disulfide cross-links on F-actin stability were only partially accounted for by the simulations of current F-actin models. Thus, our results present evidence for the high level of conformational plasticity in the interprotomer space and document the link between D-loop interactions and F-actin stability.  相似文献   

10.
The crystal structure of the F-actin binding domain 2 of severin, the gelsolin homologue from Dictyostelium discoideum, has been determined by multiple isomorphous replacement and refined to 1.75 A resolution. The structure reveals an alpha-helix-beta-sheet sandwich similar to the domains of gelsolin and villin, and contains two cation-binding sites, as observed in other domain 1 and domain 2 homologues. Comparison of the structures of several gelsolin family domains has identified residues that may mediate F-actin binding in gelsolin domain 2 homologues. To assess the involvement of these residues in F-actin binding, three mutants of human gelsolin domain 2 were assayed for F-actin binding activity and thermodynamic stability. Two of the mutants, RRV168AAA and RLK210AAA, demonstrated a lowered affinity for F-actin, indicating a role for those residues in filament binding. Using both structural and biochemical data, we have constructed a model of the gelsolin domain 1-domain 2-F-actin complex. This model highlights a number of interactions that may serve as positive and negative determinants of filament end- and side-binding.  相似文献   

11.
The first 107 residues of Fragment C of human serum albumin have been sequenced and two positions at which affinity labels block the indole site determined. Histidine 23 is the position of blockage by bromoacetyl-L-tryptophan and lysine 67 is the position of blockage by 5-dimethylaminonaphthalene-1-sulfonyl chloride and probably pyridoxal-5'-phosphate. The presence of an indole ligand at the binding site markedly reduces incorporation of the label into the above lysyl residue, and in the case of 5-dimethylaminonaphthalene-1-sulfonyl chloride, increases incorporation into three other positions, lysine residues 13, 39, and 84. It is concluded that binding of the indole ligand on the site brings about conformational changes in the albumin structure exposing new reactive positions for 5-dimethylaminonaphthalene-1-sulfonyl chloride. There is a large accumulation of basic and hydrophobic residues and no glycine, serine, threonine, valine, aspartate, or cysteine residues in the sequence 10 to 43. Lysine 71 has been identified by amino acid analyses and sequence studies as the position acetylated by acetylsalicylic acid (Hawkins, D. R., Pinckard, N., Crawford, C. P., and Farr, R. S. J. Clin. Invest. (1969) 48, 536), establishing the structural relationships of two major ligand binding sites on albumin. The lone tryptophan is at position 86. Evidence indicates that within residues 1 to 86 of Fragment C and within residues of the A-Phe fragment (Mr equals approximately 10,000), the latter known to be adjacent to Fragment C in the whole albumin structure, exists the major binding sites of all ligands for human serum albumin.  相似文献   

12.
Several metmyoglobins (red kangaroo, horse and sperm whale), containing different numbers of tyrosines, but with invariant tryptophan residues (Trp-7, Trp-14), exhibit intrinsic fluorescence when studied by steady-state front-face fluorometry. The increasing tyrosine content of these myoglobins correlates with a shift in emission maximum to shorter wavelengths with excitation at 280 nm: red kangaroo (Tyr-146) emission maximum 335 nm; horse (Tyr-103, -146) emission maximum 333 nm; sperm whale (Tyr-103, -146, -151) emission maximum 331 nm. Since 280 nm excites both tyrosine and tryptophan, this strongly suggests that tyrosine emission is not completely quenched but also contributes to this fluorescence emission. Upon titration to pH 12.5, there is a reversible shift of the emission maximum to longer wavelengths with an increase greater than 2-fold in fluorescence intensity. With excitation at 305 nm, a tyrosinate-like emission is detected at a pH greater than 12. These studies show that: (1) metmyoglobins, Class B proteins containing both tyrosine and tryptophan residues, exhibit intrinsic fluorescence; (2) tyrosine residues also contribute to the observed steady-state fluorescence emission when excited by light at 280 nm; (3) the ionization of Tyr-146 is likely coupled to protein unfolding.  相似文献   

13.
The properties of some models of the actin filament are compared with those of the thin filament in muscle. The greater stiffness of thin filaments ex vivo with respect to F-actin in vitro is attributed to the effect of both protein osmotic pressure and the attached cross-bridges. By comparing the stiffness of thin filaments in vitro and in isometric and rigor muscles the stiffness of thin filaments in relaxed muscle is computed. The upper limit of thin filament stretching is deduced to approach approximately 10 nm microm(-1). It is also calculated that, on stretching by 2.02 nm of the fully non-overlapped thin filament or by 1.59 nm of the thin filament on isometric contraction, the energy released on the hydrolysis of one molecule of ATP is fully used up.  相似文献   

14.
The Holmes F-actin model predicts a polymerization-dependent conformation change of a subdomain 3/4 loop with a hydrophobic tip (residues 266-269), allowing interaction with a hydrophobic surface on the opposing strand of the filament producing filament stabilization. We introduced cysteines in place of Val(266), Leu(267), and Leu(269) in yeast actin to allow attachment of pyrene maleimide. Pyrene at each of these positions produced differing fluorescence spectra in G-actin. Polymerization decreased the fluorescence for the 266 and 267 probes and increased that for the 269 probe. The direction of the fluorescence change was mirrored with a smaller and less hydrophobic probe, acrylodan, when attached to 266 or 269. Following polymerization, increased acrylamide quenching was observed for pyrene at 266 or 267 but not 269. The 267 probe was the least accessible of the three in G- and F-actin. F-actin quenching was biphasic for the 265, 266, and 269 but not 267 probes, suggesting that in F-actin, the pyrene samples multiple environments. Finally, in F-actin the probe at 266 interacts with one at Cys(374) on a monomer in the opposing strand, producing a pyrene excimer band. These results indicate a polymerization-dependent movement of the subdomain 3/4 loop partially consistent with Holmes' model.  相似文献   

15.
Glucosylceramide-based glycosphingolipids have been previously demonstrated to regulate negatively the formation of inositol 1,4,5-trisphosphate by phospholipase C-gamma1. In the present study, the depletion of endogenous glucosylceramide by D-t-EtDO-P4 in cultured ECV304 cells induced autophosphorylation of Src kinase at tyrosine residue 418 within the catalytic loop and dephosphorylation of Src kinase at tyrosine residues 529 within the carboxyl-terminal regulatory region. Phosphotransferase activities of Src kinase were also induced in the glucosylceramide-depleted cells. c-Src kinase activity and phosphorylations at Src Tyr-418 and epidermal growth factor (EGF) receptor Tyr-1068 were significantly enhanced by bradykinin in response to 100 nm D-t-EtDO-P4 compared with control cells. The phosphorylation and dephosphorylation on Tyr-418 and Tyr-529 residues of c-Src were reversed by treatment of 4-amino-5-(4-chlorophenyl)-7-t-butyl(pyrazolo)[3,4-d]pyrimidine (PP2), an inhibitor of Src kinase, in control cells. Glucosylceramide-depleted cells resisted treatment with PP2, and both phosphorylation of Tyr-418 and dephosphorylation of Tyr-529 induced by depletion of glucosylceramide were maintained. Compared with untreated cells, tyrosine phosphorylation of phospholipase C-gamma1 was enhanced by EGF stimulation in glucosylceramide-depleted cells, associated with enhanced tyrosine phosphorylation of the EGF receptor at Tyr-1068 and Tyr-1086 stimulated by EGF. The Src inhibitor, PP2, significantly blocked EGF-induced tyrosine phosphorylation of phospholipase C-gamma1 in control cells, whereas in glucosylceramide-depleted cells, suppression of Src kinase activity by PP2 toward EGF-induced tyrosine phosphorylation of phospholipase C-gamma1 was less significant. Thus the activation of Src kinase by depletion of glucosylceramide-based glycosphingolipids in cultured ECV304 cells is a critical up-stream event in the activation of phospholipase C-gamma1.  相似文献   

16.
A multi-site, steady-state Förster resonance energy transfer (FRET) approach was used to quantify Ca2+-induced changes in proximity between donor loci on human cardiac troponin I (cTnI), and acceptor loci on human cardiac tropomyosin (cTm) and F-actin within functional thin filaments. A fluorescent donor probe was introduced to unique and key cysteine residues on the C- and N-termini of cTnI. A FRET acceptor probe was introduced to one of three sites located on the inner or outer domain of F-actin, namely Cys-374 and the phalloidin-binding site on F-actin, and Cys-190 of cTm. Unlike earlier FRET analyses of protein dynamics within the thin filament, this study considered the effects of non-random distribution of dipoles for the donor and acceptor probes. The major conclusion drawn from this study is that Ca2+ and myosin S1-binding to the thin filament results in movement of the C-terminal domain of cTnI from the outer domain of F-actin towards the inner domain, which is associated with the myosin-binding. A hinge-linkage model is used to best-describe the finding of a Ca2+-induced movement of the C-terminus of cTnI with a stationary N-terminus. This dynamic model of the activation of the thin filament is discussed in the context of other structural and biochemical studies on normal and mutant cTnI found in hypertrophic cardiomyopathies.  相似文献   

17.
Interaction of phalloidin with chemically modified actin   总被引:3,自引:0,他引:3  
Modification of Tyr-69 with tetranitromethane impairs the polymerizability of actin in accordance with the previous report [Lehrer, S. S. and Elzinga, M. (1972) Fed. Proc. 31, 502]. Phalloidin induces this chemically modified actin to form the same characteristic helical thread-like structure as normal F-actin. The filaments bind myosin heads and activate the myosin ATPase activity as effectively as normal F-actin. When a dansyl group is introduced at the same point [Chantler, P. D. and Gratzer, W. B. (1975) Eur. J. Biochem. 60, 67-72], phalloidin still induces the polymerization. The filaments bind myosin heads and activate the myosin ATPase activity. These results indicate that Tyr-69 is not directly involved in either an actin-actin binding site or the myosin binding site on actin. Moreover, the results suggest that phalloidin binds to actin monomer in the presence of salt and its binding induces a conformational change in actin which is essential for polymerization, or that actin monomer fluctuates between in unpolymerizable and polymerizable form while phalloidin binds to actin only in the polymerizable form and its binding locks the conformation which causes the irreversible polymerization of actin. Modification of Tyr-53 with 5-diazonium-(1H)tetrazole blocks actin polymerization [Bender, N., Fasold, H., Kenmoku, A., Middelhoff, G. and Volk, K. E. (1976) Eur. J. Biochem. 64, 215-218]. Phalloidin is unable to induce the polymerization of this modified actin nor does it bind to it. Phalloidin does not induce the polymerization of the trypsin-digested actin core. These results indicate that the site at which phalloidin binds is involved in polymerization and the probable conformational change involved in polymerization may be modulated through this site.  相似文献   

18.
Three-dimensional (3-D) helical reconstructions computed from electron micrographs of negatively stained dispersed F-actin filaments invariably revealed two uninterrupted columns of mass forming the "backbone" of the double-helical filament. The contact between neighboring subunits along the thus defined two long-pitch helical strands was spatially conserved and of high mass density, while the intersubunit contact between them was of lower mass density and varied among reconstructions. In contrast, phalloidinstabilized F-actin filaments displayed higher and spatially more conserved mass density between the two long-pitch helical strands, suggesting that this bicyclic hepta-peptide toxin strengthens the intersubunit contact between the two strands. Consistent with this distinct intersubunit bonding pattern, the two long-pitch helical strands of unstabilized filaments were sometimes observed separated from each other over a distance of two to six subunits, suggesting that the intrastrand intersubunit contact is also physically stronger than the interstrand contact. The resolution of the filament reconstructions, extending to 2.5 nm axially and radially, enabled us to reproducibly "cut out" the F-actin subunit which measured 5.5 nm axially by 6.0 nm tangentially by 3.2 nm radially. The subunit is distinctly polar with a massive "base" pointing towards the "barbed" end of the filament, and a slender "tip" defining its "pointed" end (i.e., relative to the "arrowhead" pattern revealed after stoichiometric decoration of the filaments with myosin subfragment 1). Concavities running approximately parallel to the filament axis both on the inner and outer face of the subunit define a distinct cleft separating the subunit into two domains of similar size: an inner domain confined to radii less than or equal to 2.5-nm forms the uninterrupted backbone of the two long-pitch helical strands, and an outer domain placed at radii of 2-5-nm protrudes radially and thus predominantly contributes to the outer part of the massive base. Quantitative evaluation of successive crossover spacings along individual F-actin filaments revealed the deviations from the mean repeat to be compensatory, i.e., short crossovers frequently followed long ones and vice versa. The variable crossover spacings and diameter of the F-actin filament together with the local unraveling of the two long-pitch helical strands are explained in terms of varying amounts of compensatory "lateral slipping" of the two strands past each other roughly perpendicular to the filament axis. This intrinsic disorder of the actin filament may enable the actin moiety to play a more active role in actin-myosin-based force generation than merely act as a rigid passive cable as has hitherto been assumed.  相似文献   

19.
I reported previously (Higashi-Fujime, S., 1982, Cold Spring Harbor Symp. Quant. Biol., 46:69-75) that active movements of fibrils composed of F-actin and myosin filaments occurred after superprecipitation in the presence of ATP at low ionic strengths. When the concentration of MgCl2 in the medium used in the above experiment was raised to 20-26 mM, bundles of F-actin filaments, in addition to large precipitates, were formed spontaneously both during and after superprecipitation. Along these bundles, many myosin filaments were observed to slide unidirectionally and successively through the bundle, from one end to the other. The sliding of myosin filaments continued for approximately 1 h at room temperature at a mean rate of 6.0 micron/s, as long as ATP remained in the medium. By electron microscopy, it was found that most F-actin filaments decorated with heavy meromyosin pointed to the same direction in the bundle. Myosin filaments moved actively not only along the F-actin bundle but also in the medium. Such movement probably occurred along F-actin filaments that did not form the bundle but were dispersed in the medium, although dispersed F-actin filaments were not visible under the microscope. In this case, myosin filament could have moved in a reverse direction, changing from one F-actin filament to the other. These results suggested that the direction of movement of myosin filament, which has a bipolar structure and the potentiality to move in both directions, was determined by the polarity of F-actin filament in action.  相似文献   

20.
To define the actin-binding site within the NH2-terminal domain (residues 1-245) of chick smooth muscle alpha-actinin, we expressed a series of alpha-actinin deletion mutants in monkey Cos cells. Mutant alpha-actinins in which residues 2-19, 217-242, and 196-242 were deleted still retained the ability to target to actin filaments and filament ends, suggesting that the actin-binding site is located within residues 20-195. When a truncated alpha-actinin (residues 1-290) was expressed in Cos cells, the protein localized exclusively to filament ends. This activity was retained by a deletion mutant lacking residues 196-242, confirming that these are not essential for actin binding. The actin-binding site in alpha-actinin was further defined by expressing both wild-type and mutant actin-binding domains as fusion proteins in E. coli. Analysis of the ability of such proteins to bind to F-actin in vitro showed that the binding site was located between residues 108 and 189. Using both in vivo and in vitro assays, we have also shown that the sequence KTFT, which is conserved in several members of the alpha-actinin family of actin-binding proteins (residues 36-39 in the chick smooth muscle protein) is not essential for actin binding. Finally, we have established that the NH2-terminal domain of dystrophin is functionally as well as structurally homologous to that in alpha-actinin. Thus, a chimeric protein containing the NH2-terminal region of dystrophin (residues 1-233) fused to alpha-actinin residues 244-888 localized to actin-containing structures when expressed in Cos cells. Furthermore, an E. coli-expressed fusion protein containing dystrophin residues 1-233 was able to bind to F-actin in vitro.  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