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.
Vertebrate skeletal muscle alpha-tropomyosin polymerizes in a head-to-tail manner and binds cooperatively to actin. It has been postulated that the cooperative actin binding is governed by the strength of the head-to-tail interaction. In order to know the relationship between the head-to-tail affinity and actin binding, we studied the properties of tropomyosin variants with single residue substitutions at serine-283, the penultimate residue at the carboxyl terminus that is involved in the head-to-tail interaction. It has been shown that the phosphorylation of serine-283 strengthens the head-to-tail interaction. Viscometry was employed to compare the head-to-tail affinity of tropomyosin variants. Variant S283E showed higher viscosity whereas variant S283K showed lower viscosity compared with the wild type non-phosphorylated alpha-tropomyosin. The results confirm the idea that the interaction is sensitive to the ionic properties of residue 283. The strength of the head-to-tail interaction was assessed directly by sedimentation equilibrium using two pairs of tropomyosin variants designed so that only dimeric interactions were allowed within each pair. From one pair of variants with serine-283, the association constant was determined to be 2.6 x 10(4) M(-1) (SD =1.0 x 10(4)), whereas for the second pair with glutamate-283, the affinity was 3.9 x 10(4) M(-1) (SD =1.6 x 10(4)), slightly stronger than the former, consistent with the results of viscometry. The results indicate that the head-to-tail association is weak as previously implicated from light scattering measurements. Cosedimentation was employed to measure the cooperative actin binding of tropomyosin variants. Although previous results indicated the phosphorylation has no significant influence on the actin affinity, variant S283E shows a lower affinity compared with the control. Variants S283K and S283A show even lower affinities to actin, although these species bind to actin more cooperatively than does variant S283E. The results indicate that the affinity of the head-to-tail interaction between adjacent tropomyosin molecules is weak, and is substantially influenced by an extra charge at residue 283. On the other hand, the interaction with actin, the affinity and the cooperativity in actin binding, is dependent on amino acid residues at 283 and is not simply correlated with the strength of the head-to-tail interaction between Tm molecules in solution.  相似文献   

2.
J M Mo  M E Holtzer  A Holtzer 《Biopolymers》1990,30(9-10):921-927
Nonpolymerizable tropomyosin (NPTm) is found to unfold thermally at high ionic strength almost exactly as the parent protein, but it does not aggregate at low ionic strength. Thus, NPTm can be used as a tropomyosin surrogate whose coiled-coil structural stability can be probed by varying the ionic strength. Studies of NPTm by CD show that increasing ionic strength stabilizes the coiled-coil structure. CD spectra over a wide range of helix content, obtained by varying either temperature or ionic strength, show an isodichroic point at 203 nm, suggesting a local, residue-level, two-state model. At given temperature, such a local helix in equilibrium random equilibrium suggests ln [phi h/(1-phi h)] = A1 + A2In, wherein phi h is the fraction helix, and A1, A2, and n are constants. In the low ionic strength region, theoretical limiting laws for ionic strength mediated charge-charge, dipole-dipole, and apolar-apolar (salting out) interactions give, respectively, n = 0.5, 1.0, and 1.0. Our experimental values for 40 degrees C, where the data span a wide range of helix content, show n = 1.0, suggesting that ionic strength stabilizes either by reducing dipole-dipole repulsions or by enhancing hydrophobic interactions, both probably interhelix in nature. Two segments of tropomyosin, 11Tm127 and 142Tm281, neither of which aggregate at low ionic strength, give results similar to those for NPTm, i.e., n = 0.96 and 0.84, respectively.  相似文献   

3.
Tropomyosin is a 284 residue dimeric coiled-coil protein that interacts in a head-to-tail manner to form linear filaments at low ionic strengths. Polymerization is related to tropomyosin's ability to bind actin, and both properties depend on intact N- and C-termini as well as alpha-amino acetylation of the N-terminus of the muscle protein. Nalpha-acetylation can be mimicked by an N-terminal Ala-Ser fusion in recombinant tropomyosin (ASTm) produced in Escherichia coli. Here we show that a recombinant tropomyosin fragment, corresponding to the protein's first 260 residues plus an Ala-Ser fusion [ASTm(1-260)], polymerizes to a much greater extent than the corresponding full-length recombinant protein, despite the absence of the C-terminal 24 amino acids. This polymerization is sensitive to ionic strength and is greatly reduced by the removal of the N-terminal Ala-Ser fusion [nfTm(1-260)]. CD studies show that nonpolymerizable tropomyosin fragments, which terminate at position 260 [Tm(167-260) and Tm(143-260)], as well as Tm(220-284), are able to interact with ASTm(1-142), a nonpolymerizable N-terminal fragment, and that the head-to-tail interactions observed for these fragment pairs are accompanied by a significant degree of folding of the C-terminal tropomyosin fragment. These results suggest that the new C-terminus, created by the deletion, polymerizes in a manner similar to the full-length protein. Head-to-tail binding for fragments terminating at position 260 may be explained by the presence of a greater concentration of negatively charged residues, while, at the same time, maintaining a conserved pattern of charged and hydrophobic residues found in polymerizable tropomyosins from a variety of sources.  相似文献   

4.
We have used x-ray diffraction to study the interactions between myelin membranes in the sciatic nerve (PNS) and optic nerve (CNS) as a function of pH (2-10) and ionic strength (0-0.18). The period of myelin was found to change in a systematic manner with pH and ionic strength. PNS periods ranged from 165 to 250 A or more, while CNS periods ranged from 150 to 230 A. The native periods were observed only near physiological ionic strength at neutral or alkaline pH. The smallest periods were observed in the pH range 2.5-4 for PNS myelin and pH 2.5-5 for CNS myelin. The minimum period was also observed for PNS myelin after prolonged incubation in distilled water. At pH 4, within these acidic pH ranges, myelin period increased slightly with ionic strength; however, above these ranges, the period increased with pH and decreased with ionic strength. Electron density profiles calculated at different pH and ionic strength showed that the major structural alteration underlying the changes in period was in the width of the aqueous space at the extracellular apposition of membranes; the width of the cytoplasmic space was virtually constant. Assuming that the equilibrium myelin periods are determined by a balance of nonspecific forces/i.e., the electrostatic repulsion force and the van der Walls attractive force, as well as the short-range repulsion force (hydration force, or steric stabilization), then values in the period-dependency curve can be used to define the isoelectric pH and exclusion length of the membrane. The exclusion length, which is related to the minimum period at isoelectric pH, was used to calculate the electrostatic repulsion force given the other forces. The electrostatic repulsion was then used to calculate the surface potential, which in turn was used to calculate the surface charge density (at different pH and ionic strength). We found the negative surface charge increases with pH at constant ionic strength and with ionic strength at constant pH. We suggest that the former is due to deprotonation of the ionizable groups on the surface while the latter is due to ion binding. Interpretation of our data in terms of the chemical composition of myelin is given in the accompanying paper (Inouye and Kirschner, 1988). We also calculated the total potential energy functions for the different equilibrium periods and found that the energy minima became shallower and broader with increasing membrane separation. Finally, it was difficult to account directly for certain structural transitions from a balance of nonspecific forces.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
Tropomyosin (Tm) is a dimeric coiled-coil protein, composed of 284 amino acids (410 A), that forms linear homopolymers through head-to-tail interactions at low ionic strength. The head-to-tail complex involves the overlap of approximately nine N-terminal residues of one molecule with nine C-terminal residues of another Tm molecule. In this study, we investigate the influence of 2,2,2-trifluoroethanol (TFE) and glycerol on the stability of recombinant Tm fragments (ASTm1-142, Tm143-284(5OHW269)) and of the dimeric head-to-tail complex formed by the association of these two fragments. The C-terminal fragment (Tm143-284(5OHW269)) contains a 5-hydroxytryptophan (5OHW) probe at position 269 whose fluorescence is sensitive to the head-to-tail interaction and allows us to accompany titrations of Tm143-284(5OHW269) with ASTm1-142 to calculate the dissociation constant (Kd) and the interaction energy at TFE and glycerol concentrations between 0% and 15%. We observe that TFE, but not glycerol, reduces the stability of the head-to-tail complex. Thermal denaturation experiments also showed that the head-to-tail complex increases the overall conformational stability of the Tm fragments. Urea and thermal denaturation assays demonstrated that both TFE and glycerol increase the stability of the isolated N- and C-terminal fragments; however, only TFE caused a significant reduction in the cooperativity of unfolding these fragments. Our results show that these two cosolvents stabilize the structures of individual Tm fragments in different manners and that these differences may be related to their opposing effects on head-to-tail complex formation.  相似文献   

6.
G L Lilley  L W Fung 《Life sciences》1987,41(22):2429-2436
The spin-label electron paramagnetic resonance (EPR) technique has been used to study the interaction between human hemoglobin and erythrocyte membranes as a function of temperature and ionic strength. We show, for the first time, experimental evidence for the existence of the interaction at physiological pH, ionic strength and temperature. In addition to the pH dependence that we have previously reported, the interactions are also temperature and ionic strength dependent. Using a simple two-state equilibrium model to analyze the EPR data, we obtain an equilibrium dissociation constant of about 8.1 +/- 5.6 X 10(-5) M for hemoglobin-membrane systems in 5 mM phosphate with 150 mM NaCl at pH 7.4 and 37 degrees C.  相似文献   

7.
Tropomyosin isoforms of the low Mr class were isolated from chicken intestinal epithelium and brain, and their physical and functional properties were characterized. Tropomyosin from each tissue contains four distinct polypeptides, all of about 32,000 daltons. In two-dimensional gels, brain tropomyosin contains two major and two minor polypeptides; the major epithelium isoforms coelectrophorese with the two minor brain isoforms. Conversely, only small amounts of the major brain isoforms are detected in the epithelium. Actin-binding properties of brain tropomyosin isoforms are distinct from those of the intestinal epithelium. At 2.5 mM MgCl2 and physiological ionic strength, the intestinal epithelial tropomyosin binds to filamentous actin with an apparent Ka of 8 X 10(6) M-1 whereas brain tropomyosin has an apparent Ka of 8 X 10(5) M-1. Tropomyosin from either tissue binds actin cooperatively with a Hill coefficient of 2.3 for intestinal epithelial cell and 1.95 for brain tropomyosin. Isoforms from both tissues exhibit reduced head-to-tail polymerizability as compared to muscle tropomyosin. The actin-binding properties of intestinal epithelial cell tropomyosin are therefore similar to those of the muscle tropomyosins even though the isoforms have lower molecular weight, a paracrystal structure, and reduced head-to-tail polymerizability typical of the other nonmuscle tropomyosins. These results indicate that a heterogeneity of functional properties may be expressed among the low Mr tropomyosin isoforms.  相似文献   

8.
The equilibrium trajectory of the axis of a rod subject to an externally imposed curved potential energy trough tends to conform to the shape of the curved trough, but also tends to be straight because of elastic resistance to bending. The actual path of the axis is a balance between the two extremes. We consider a potential energy trough centered along a circular arc of radius R. For a rod of small length compared to R, we show that the axis at equilibrium forms an arc of a circle of radius greater than R. The value of the radius of the axial path depends on the relative values of the Hooke's Law bending constant for the rod and the depth and width of the trough. Motivation for the calculation is provided by nucleosomal DNA, which conforms to the surface of a roughly cylindrical histone core at physiological ionic strength, but is observed to unwind into a partially extended conformation at very low ionic strength. We suggest that the rigidity to bending of short DNA segments becomes sufficiently great at low ionic strength to overcome attractive interactions with the histone surface. Alternately, of course, if during the cell cycle mutually attractive forces between DNA and histone core are weakened at constant ionic strength, the same type of unfolding would be expected to occur as the strength of the DNA-histone contacts drops below the level required to overcome elastic resistance to bending of the DNA rod.  相似文献   

9.
A semi–quantitative theory is developed to explain the nonspecific binding of proteins to substituted affinity chromatography supports due to electrostatic and hydrophobic interactions. The equilibrium constant for the absorption of an enzyme to a solid support, and the rate of desorption of the enzyme are studied as functions of ionic strength. Experimental measurements were taken of the adsorption equilibrium constant and rate of desorption of E. coli β–galactosidase on Sepharose 4B substituted with 3, 3,-diaminodipropylamine in batch systems. It was found that the enzyme adsorption exhibits a hysteresis effect as the ionic strength is increased and then decreased. Furthermore, the adsorption of theenzyme becomes more reversible at the lower ionic strengths, while at the higher ionic strengths it is essentially irreversible. Using the measured equilibrium constants, and knowing the region of ionic strength where the adsorption becomes reversible, we were able to predict the desorption of enzyme in a continuous stirred tank as a function of time when a decreasing linear gradient of ionic strength was introduced into a slurry. It was found that the presence of another protein, hemoglobin, does not affect these results, and therefore can be separated from the enzyme.  相似文献   

10.
Heavy meromyosin from scallop (scHMM) striated muscle is regulated by calcium binding to the essential light chain. The regulation can be modeled with a calcium-dependent equilibrium between on and off scHMM conformations. The observed rate constant for mant-ADP dissociation from scHMM is calcium dependent, and we show here that it can be used to define the equilibrium constant (K(eq)) between on and off conformations. The data show that K(eq) is markedly ionic strength dependent, with high salt (>/=200 mM) abolishing the off state even in the absence of calcium and low salt (<50 mM) favoring the off state even in the presence of calcium. Debye-Huckel plots of the equilibrium constant (K(eq)) for the on and off forms gave parallel slopes (5.94 +/- 0.33 and 6.36 +/- 0.17 M(-0.5)) in the presence and absence of calcium. The presence of an equilibrium mixture of two conformations was confirmed by sedimentation data and the effects of ADP, calcium and ionic strength were in qualitative agreement. Thus scHMM exists in two conformations that can be distinguished in sedimentation profiles and by the rate of release of mant-ADP. Increasing salt concentrations biases the system toward the on state, suggesting a role for ionic interactions in stabilizing the off state.  相似文献   

11.
《Biophysical journal》2019,116(12):2275-2284
The initial binding of tropomyosin onto actin filaments and then its polymerization into continuous cables on the filament surface must be precisely tuned to overall thin-filament structure, function, and performance. Low-affinity interaction of tropomyosin with actin has to be sufficiently strong to localize the tropomyosin on actin, yet not so tight that regulatory movement on filaments is curtailed. Likewise, head-to-tail association of tropomyosin molecules must be favorable enough to promote tropomyosin cable formation but not so tenacious that polymerization precedes filament binding. Arguably, little molecular detail on early tropomyosin binding steps has been revealed since Wegner’s seminal studies on filament assembly almost 40 years ago. Thus, interpretation of mutation-based actin-tropomyosin binding anomalies leading to cardiomyopathies cannot be described fully. In vitro, tropomyosin binding is masked by explosive tropomyosin polymerization once cable formation is initiated on actin filaments. In contrast, in silico analysis, characterizing molecular dynamics simulations of single wild-type and mutant tropomyosin molecules on F-actin, is not complicated by tropomyosin polymerization at all. In fact, molecular dynamics performed here demonstrates that a midpiece tropomyosin domain is essential for normal actin-tropomyosin interaction and that this interaction is strictly conserved in a number of tropomyosin mutant species. Elsewhere along these mutant molecules, twisting and bending corrupts the tropomyosin superhelices as they “lose their grip” on F-actin. We propose that residual interactions displayed by these mutant tropomyosin structures with actin mimic ones that occur in early stages of thin-filament generation, as if the mutants are recapitulating the assembly process but in reverse. We conclude therefore that an initial binding step in tropomyosin assembly onto actin involves interaction of the essential centrally located domain.  相似文献   

12.
Alberty RA 《Biochemistry》2004,43(30):9840-9845
Recent thermodynamic measurements have made it possible to calculate the apparent equilibrium constants of the ribonucleoside diphosphate reductase reaction and the ribonucleoside triphosphate reductase reaction with various reducing agents. Third law heat capacity measurements on crystals of d-ribose and other calorimetric measurements make it possible to calculate Delta(f)G degrees for D-ribose and two species of D-ribose 5-phosphate. The experimental value of the apparent equilibrium constant K' for the deoxyribose-phosphate aldolase reaction makes it possible to calculate the standard Gibbs energies of formation Delta(f)G degrees for two protonation states of 2'-deoxy-D-ribose 5-phosphate. This shows that Delta(f)G degrees (2'-deoxy-D-ribose 5-phosphate(2)(-)) - Delta(f)G degrees (D-ribose 5-phosphate(2)(-)) = 147.86 kJ mol(-1) at 298.15 K and zero ionic strength in dilute aqueous solutions. This difference between reduced and oxidized forms is expected to apply to D-ribose, D-ribose 1-phosphate, ribonucleosides, and ribonucleotides in general. This expectation is supported by two other enzyme-catalyzed reactions for which apparent equilibrium constants have been determined. The availability of Delta(f)G degrees values for the species of 2'-deoxy-D-ribose and its derivatives makes it possible to calculate standard transformed Gibbs energies of formation of these reactants, apparent equilibrium constants for their reactions, changes in the binding of hydrogen ions in these reactions, and standard apparent reduction potentials of the half reactions involved as a function of pH and ionic strength at 298.15 K. The apparent equilibrium constant for ADP + thioredoxin(red) = 2'-deoxyADP + H(2)O + thioredoxin(ox) is 1.4 x 10(11) at 298.15 K, pH 7, and 0.25 M ionic strength.  相似文献   

13.
Human erythrocytes contain a Mr 43,000 tropomyosin-binding protein that is unrelated to actin and that has been proposed to play a role in modulating the association of tropomyosin with spectrin-actin complexes based on its stoichiometry in the membrane skeleton of one Mr 43,000 monomer per short actin filament (Fowler, V. M. 1987. J. Biol. Chem. 262:12792-12800). Here, we describe an improved procedure to purify milligram quantities to 98% homogeneity and we show that this protein inhibits tropomyosin binding to actin by a novel mechanism. We have named this protein tropomodulin. Unlike other proteins that inhibit tropomyosin-actin interactions, tropomodulin itself does not bind to F-actin. EM of rotary-shadowed tropomodulin-tropomyosin complexes reveal that tropomodulin (14.5 +/- 2.4 nm [SD] in diameter) binds to one of the ends of the rod-like tropomyosin molecules (33 nm long). In agreement with this observation, Dixon plots of inhibition curves demonstrate that tropomodulin is a non-competitive inhibitor of tropomyosin binding to F-actin (Ki = 0.7 microM). Hill plots of the binding of the tropomodulin-tropomyosin complex to actin indicate that binding does not exhibit any positive cooperativity (n = 0.9), in contrast to tropomyosin (n = 1.9), and that the apparent affinity of the complex for actin is reduced 20-fold with respect to that of tropomyosin. These results suggest that binding of tropomodulin to tropomyosin may block the ability of tropomyosin to self-associate in a head-to-tail fashion along the actin filament, thereby weakening its binding to actin. Antibodies to tropomodulin cross-react strongly with striated muscle troponin I (but not with troponin T) as well as with a nontroponin Mr 43,000 polypeptide in muscle and in other nonerythroid cells and tissues, including brain, lens, neutrophils, and endothelial cells. Thus, erythrocyte tropomodulin may be one member of a family of tropomyosin-binding proteins that function to regulate tropomyosin-actin interactions in non-muscle cells and tissues.  相似文献   

14.
Smooth muscle contraction is controlled in part by the state of phosphorylation of myosin. A recently discovered actin and calmodulin-binding protein, named caldesmon, may also be involved in regulation of smooth muscle contraction. Caldesmon cross-links actin filaments and also inhibits actin-activated ATP hydrolysis by myosin, particularly in the presence of tropomyosin. We have studied the effect of caldesmon on the rate of hydrolysis of ATP by skeletal muscle myosin subfragment-1, a system in which phosphorylation of the myosin is not important in regulation. Caldesmon is a very effective inhibitor of ATP hydrolysis giving up to 95% inhibition. At low ionic strength (approximately 20 mM) this effect does not require smooth muscle tropomyosin, whereas at high ionic strength (approximately 120 mM) tropomyosin enhances the inhibitory activity of caldesmon at low caldesmon concentrations. Cross-linking of actin is not essential for inhibition of ATP hydrolysis to occur since at high ionic strength there is very little cross-linking as determined by a low speed sedimentation assay. Under all conditions examined, the decrease in the rate of ATP hydrolysis is accompanied by a decrease in the binding of myosin subfragment-1 to actin. Furthermore, caldesmon weakens the equilibrium binding of myosin subfragment-1 to actin in the presence of pyrophosphate. We conclude that caldesmon has a general weakening effect on the binding of skeletal muscle myosin subfragment-1 to actin and that this weakening in binding may be responsible for inhibition of ATP hydrolysis.  相似文献   

15.
The equilibrium trajectory of the axis of a rod subject to an externally imposed curved potential energy trough tends to conform to the shape of the curved trough, but also tends to be straight because of elastic resistance to bending. The actual path of the axis is a balance between the two extremes. We consider a potential energy trough centered along a circular arc of radiusR. For a rod of small length compared toR, we show that the axis at equilibrium forms an arc of a circle of radius greater thanR. The value of the radius of the axial path depends on the relative values of the Hooke’s Law bending constant for the rod and the depth and width of the trough. Motivation for the calculation is provided by nucleosomal DNA, which conforms to the surface of a roughtly cylindrical histone core at physiological ionic strength, but is observed to unwind into a partially extended conformation at very low ionic strength. We suggest that the rigidity to bending of short DNA segments becomes sufficiently great at low ionic strength to overcome attractive interactions with the histone surface. Alternately, of course, if during the cell cycle mutually attractive forces between DNA and histone core are weakened at constant ionic strength, the same type of unfolding would be expected to occur as the strength of the DNA-histone contacts drops below the level required to overcome elastic resistance to bending of the DNA rod.  相似文献   

16.
The best way to store data on apparent equilibrium constants for enzyme-catalyzed reactions is to calculate the standard Gibbs energies of formation of the species involved at 298.15 K and zero ionic strength so that equilibrium constants can be calculated at the desired pH and ionic strength. These calculations are described for CoA, acetyl-CoA, oxalyl-CoA, succinyl-CoA, methylmalonyl-CoA, malyl-CoA and CoA-glutathione. The species properties are then used to calculate standard transformed Gibbs energies of formation for these reactants as functions of pH at ionic strength 0.25 M. The species data also make it possible to calculate apparent equilibrium constants of 23 enzyme-catalyzed reactions as a function of pH, including some that cannot be determined directly because they are so large.  相似文献   

17.
The role of the overlap region at the ends of tropomyosin molecules in the properties of regulated thin filaments has been investigated by substituting nonpolymerizable tropomyosin for tropomyosin in a reconstituted troponin-tropomyosin-actomyosin subfragment 1 ATPase assay system. A previous study [Heeley, Golosinka & Smillie (1987) J. Biol. Chem. 262, 9971-9978] has shown that at an ionic strength of 70 mM, troponin will induce full binding of nonpolymerizable tropomyosin to F-actin both in the presence and absence of calcium. At a myosin subfragment 1-to-actin ratio of 2:1 ([actin] = 4 microM) and an ionic strength of 50 mM, comparable levels of ATPase inhibition were observed with increasing levels of tropomyosin or the truncated derivative in the presence of troponin (-Ca2+). Large differences were noted, however, in the activation by Ca2+. Significantly lower ATPase activities were observed with nonpolymerizable tropomyosin and troponin (+Ca2+) over a range of subfragment 1-to-actin ratios from 0.25 to 2.5. The concentration of subfragment 1 required to generate ATPase activities exceeding those seen with actomyosin subfragment 1 alone under these conditions was 3-4-fold greater when nonpolymerizable tropomyosin was used. Similar effects were seen at the much lower ionic strength of 13 mM and are consistent with the reduced ATPase activity with nonpolymerizable tropomyosin observed previously [Walsh, Trueblood, Evans & Weber (1985) J. Mol. Biol. 182, 265-269] at low ionic strength and a subfragment 1-to-actin ratio of 1:100. Little cooperativity in activity as a function of subfragment 1 concentration with either intact tropomyosin or its truncated derivative was observed under the present conditions. Further studies are directed towards an understanding of these effects in terms of the two-state binding model for the attachment of myosin heads to regulated thin filaments.  相似文献   

18.
Dilated cardiomyopathy (DCM) is associated with mutations in cardiomyocyte sarcomeric proteins, including α-tropomyosin. In conjunction with troponin, tropomyosin shifts to regulate actomyosin interactions. Tropomyosin molecules overlap via tropomyosin–tropomyosin head-to-tail associations, forming a continuous strand along the thin filament. These associations are critical for propagation of tropomyosin''s reconfiguration along the thin filament and key for the cooperative switching between heart muscle contraction and relaxation. Here, we tested perturbations in tropomyosin structure, biochemistry, and function caused by the DCM-linked mutation, M8R, which is located at the overlap junction. Localized and nonlocalized structural effects of the mutation were found in tropomyosin that ultimately perturb its thin filament regulatory function. Comparison of mutant and WT α-tropomyosin was carried out using in vitro motility assays, CD, actin co-sedimentation, and molecular dynamics simulations. Regulated thin filament velocity measurements showed that the presence of M8R tropomyosin decreased calcium sensitivity and thin filament cooperativity. The co-sedimentation of actin and tropomyosin showed weakening of actin-mutant tropomyosin binding. The binding of troponin T''s N terminus to the actin-mutant tropomyosin complex was also weakened. CD and molecular dynamics indicate that the M8R mutation disrupts the four-helix bundle at the head-to-tail junction, leading to weaker tropomyosin–tropomyosin binding and weaker tropomyosin–actin binding. Molecular dynamics revealed that altered end-to-end bond formation has effects extending toward the central region of the tropomyosin molecule, which alter the azimuthal position of tropomyosin, likely disrupting the mutant thin filament response to calcium. These results demonstrate that mutation-induced alterations in tropomyosin–thin filament interactions underlie the altered regulatory phenotype and ultimately the pathogenesis of DCM.  相似文献   

19.
Phosphorylated rabbit cardiac alpha alpha-tropomyosin has been prepared either enzymatically (Montgomery, K., and Mak, A.S. (1984) J. Biol. Chem. 259, 5555-5560) or by fractionation of the phosphorylated and nonphosphorylated forms on a Mono Q column in 9 M urea, 50 mM Tris, pH 8.0. Although the phosphorylated and nonphosphorylated forms showed no difference in their F-actin binding properties, the phosphorylated protein had substantially higher viscosities at low ionic strengths, indicating a greater propensity for head-to-tail interaction. Similar measurements showed the strengthening of this interaction by whole troponin to be substantially reduced by phosphorylation even though the binding of whole troponin and troponin T to tropomyosin was demonstrated by affinity chromatography to be, if anything, strengthened by phosphorylation. In a reconstituted actin (4 microM) plus myosin subfragment 1 ATPase assay (50 mM ionic strength), significantly higher activities over a range (1 to 8 microM) of subfragment 1 concentrations were observed with phosphorylated tropomyosin compared with the nonphosphorylated protein. In the fully reconstituted system with troponin, there was no significant difference in the inhibition of ATPase in the absence of Ca2+. However, in its presence, the activities were appreciably increased with the phosphorylated tropomyosin compared to those with the nonphosphorylated form. These differences were eliminated by treatment of the phosphorylated tropomyosin with alkaline phosphatase. This is the first demonstration of an effect of phosphorylation on the functional properties of tropomyosin.  相似文献   

20.
We have expressed two variants of chicken striated muscle alpha-tropomyosin in Escherichia coli: fusion tropomyosin containing 80 amino acids of a non-structural influenza virus protein (NS1) on the amino terminus and a non-fusion tropomyosin which is a variant because the amino-terminal methionine is not acetylated (unacetylated tropomyosin). From our analysis of purified proteins in vitro we suggest that the amino-terminal region, which is highly conserved in muscle tropomyosins, is crucial for all aspects of tropomyosin function. Both forms are altered in tropomyosin activity: neither shows head-to-tail polymerization, with or without troponin. Unacetylated tropomyosin binds weakly to actin, but in the presence of troponin it binds well and can regulate the actomyosin ATPase. Fusion tropomyosin binds well to actin, but binding of troponin is calcium-sensitive and it does not confer effective calcium sensitivity on the actomyosin ATPase. Our results indicate that the local charge at the amino terminus is critical for actin binding but that normal head-to-tail association is not required. The properties of fusion tropomyosin-troponin interaction are indicative of impaired troponin T binding to tropomyosin and provide evidence for its binding to the amino terminus of tropomyosin.  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