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.
Millisecond internal dynamics of native and inactivated actin from rabbit skeletal muscle was examined using room temperature phosphorescence. Inactivated actin was prepared by incubation of G-actin at 70 degrees C, by treatment with 4 M urea or 1.5 M guanidinium hydrochloride, renaturation from fully unfolded state or by Ca2+ ion removal. It was shown that inactivation of actin, irrespective of the denaturation procedure applied, leads to a sharp decrease of millisecond fluctuations of the protein structure. Restriction of the slow intramolecular mobility in inactivated actin can result from changes of the protein conformation and/or specific association of macromolecules.  相似文献   

2.
Thymosin beta 4 (T beta 4), a 5-kD peptide which binds G-actin and inhibits its polymerization (Safer, D., M. Elzinga, and V. T. Nachmias. 1991. J. Biol. Chem. 266:4029-4032), appears to be the major G-actin sequestering protein in human PMNs. In support of a previous study by Hannappel, E., and M. Van Kampen (1987. J. Chromatography. 397:279-285), we find that T beta 4 is an abundant peptide in these cells. By reverse phase HPLC of perchloric acid supernatants, human PMNs contain approximately 169 fg/cell +/- 90 fg/cell (SD), corresponding to a cytoplasmic concentration of approximately 149 +/- 80.5 microM. On non-denaturing polyacrylamide gels, a large fraction of G-actin in supernatants prepared from resting PMNs has a mobility similar to the G-actin/T beta 4 complex. Chemoattractant stimulation of PMNs results in a decrease in this G-actin/T beta 4 complex. To determine whether chemoattractant induced actin polymerization results from an inactivation of T beta 4, the G-actin sequestering activity of supernatants prepared from resting and chemoattractant stimulated cells was measured by comparing the rates of pyrenyl-actin polymerization from filament pointed ends. Pyrenyl actin polymerization was inhibited to a greater extent in supernatants from stimulated cells and these results are qualitatively consistent with T beta 4 being released as G-actin polymerizes, with no chemoattractant-induced change in its affinity for G-actin. The kinetics of bovine spleen T beta 4 binding to muscle pyrenyl G-actin are sufficiently rapid to accommodate the rapid changes in actin polymerization and depolymerization observed in vivo in response to chemoattractant addition and removal.  相似文献   

3.
On the elastic properties of tetramethylrhodamine F-actin   总被引:2,自引:0,他引:2  
(Iodoacetamido)tetramethylrhodamine disrupts F-actin. At the 1:1 fluorophore to actin (as monomer) ratio approximately 80% of the protein becomes non-sedimentable. The fluorescent, non-sedimentable actin copolymerizes with G-actin to yield fluorescent filaments. The tensile strength of these filaments changes with the ratio of the fluorescent non-sedimentable actin to the G-actin, being 1.6 pN, 2.9 pN and 3.6 pN at the 1/4, 2/3 and 1/1 ratios, respectively. These tensile strengths are approximately two orders of magnitude lower than those obtained by decoration of F-actin with phalloidin.  相似文献   

4.
We previously observed that the neurotransmitter 5-hydroxytryptamine (5-HT, serotonin) binds with high- and low-affinity interactions to an actin-like protein prepared from rat brain synaptosomes. In this study, we examined its binding to highly purified actin obtained from rabbit skeletal muscle. Monomeric G-actin bound serotonin with high and low affinities, exhibiting equilibrium dissociation constants (KD values) of 5 X 10(-5) M and 4 X 10(-3) M, respectively. The serotonin binding site on actin was distinct from those sites previously characterized for divalent cations, nucleotides, and cytochalasin alkaloids. The binding of serotonin (1 microM) to G-actin was increased as much as 26-fold by divalent cations. Potassium iodine (KI) increased the affinity of G-actin for serotonin, KD values for this binding being 3 X 10(-7) M and X 10(-5) M. Serotonin bound with even higher affinity to polymerized F-actin, with KD values of 2 X 10(-8) M and 2 X 10(-5) M. However, the total number of binding sites on F-actin was only about 4% of the number of G-actin. The binding of serotonin (0.1 microM) to G-actin could be inhibited by phenothiazines (1 microM) or reserpine (10 microM), but not by classical antagonists of serotonin receptors or by drugs that release serotonin or inhibit its uptake. The binding of serotonin to actin in vivo may participate in a contractile process related to neurotransmitter release.  相似文献   

5.
The red cell membrane skeletal network is constructed from actin, spectrin and protein 4.1 in a molar ratio of actin subunits/spectrin heterodimer/protein 4.1 of 2:1:1. This represents saturation of the actin filaments, since incubation with extraneous spectrin and protein 4.1 leads to no binding of additional spectrin, either to the inner surface of ghost membranes or to lipid-free membrane cytoskeletons. Partial extraction of spectrin from the membrane is accompanied by release of actin under all conditions. Regardless of the proportion of spectrin extracted, the molar ratio of spectrin dimers/actin subunits is constant at 1:2. This is not the result of release or cooperative breakdown of whole lattice junctions from the network, for the number of actin filaments, judged by capacity to nucleate polymerisation of added G-actin, remains unchanged even when as much as 60% of the total spectrin has been lost. A similar 1:2:1 stoichiometry characterises the complex formed when G-actin is allowed to polymerise in the presence of varying amounts of spectrin and protein 4.1. When this complex is treated with the depolymerising agent, 1 M guanidine hydrochloride, it breaks down into smaller units of the same stoichiometry. After cross-linking these can be recovered from a gel-filtration column. Complexes prepared starting from G-actin appear to be much more stable than those formed when spectrin and protein 4.1 are bound to F-actin.  相似文献   

6.
Oxidovanadium(IV), a cationic species (VO2+) of vanadium(IV), binds to several proteins, including actin. Upon titration with oxidovanadium(IV), approximately 100% quenching of the intrinsic fluorescence of monomeric actin purified from rabbit skeletal muscle (G-actin) was observed, with a V50 of 131 μM, whereas for the polymerized form of actin (F-actin) 75% of quenching was obtained and a V50 value of 320 μM. Stern-Volmer plots were used to estimate an oxidovanadium(IV)-actin dissociation constant, with Kd of 8.2 μM and 64.1 μM VOSO4, for G-actin and F-actin, respectively. These studies reveal the presence of a high affinity binding site for oxidovanadium(IV) in actin, producing local conformational changes near the tryptophans most accessible to water in the three-dimensional structure of actin. The actin conformational changes, also confirmed by 1H NMR, are accompanied by changes in G-actin hydrophobic surface, but not in F-actin. The 1H NMR spectra of G-actin treated with oxidovanadium(IV) clearly indicates changes in the resonances ascribed to methyl group and aliphatic regions as well as to aromatics and peptide-bond amide region. In parallel, it was verified that oxidovanadium(IV) prevents the G-actin polymerization into F-actin. In the 0-200 μM range, VOSO4 inhibits 40% of the extent of polymerization with an IC50 of 15.1 μM, whereas 500 μM VOSO4 totally suppresses actin polymerization. The data strongly suggest that oxidovanadium(IV) binds to actin at specific binding sites preventing actin polymerization. By affecting actin structure and function, oxidovanadium(IV) might be responsible for many cellular effects described for vanadium.  相似文献   

7.
The propensity to associate or aggregate is one of the characteristic properties of many nonnative proteins. The aggregation of proteins is responsible for a number of human diseases and is a significant problem in biotechnology. Despite this, little is currently known about the effect of self-association on the structural properties and conformational stability of partially folded protein molecules. G-actin is shown to form equilibrium unfolding intermediate in the vicinity of 1.5 M guanidinium chloride (GdmCl). Refolding from the GdmCl unfolded state is terminated at the stage of formation of the same intermediate state. An analogous form, known as inactivated actin, can be obtained by heat treatment, or at moderate urea concentration, or by the release of Ca(2+). In all cases actin forms specific associates comprising partially folded protein molecules. The structural properties and conformational stability of inactivated actin were studied over a wide range of protein concentrations, and it was established that the process of self-association is rather specific. We have also shown that inactivated actin, being denatured, is characterized by a relatively rigid microenvironment of aromatic residues and exhibits a considerable limitation in the internal mobility of tryptophans. This means that specific self-association can play an important structure-forming role for the partially folded protein molecules.  相似文献   

8.
ATP prevents G-actin cysteine oxidation and vanadyl formation specifically induced by decavanadate, suggesting that the oxometalate-protein interaction is affected by the nucleotide. The ATP exchange rate is increased by 2-fold due to the presence of decavanadate when compared with control actin (3.1 × 10− 3 s− 1), and an apparent dissociation constant (kdapp) of 227.4 ± 25.7 μM and 112.3 ± 8.7 μM was obtained in absence or presence of 20 μM V10, respectively. Moreover, concentrations as low as 50 μM of decameric vanadate species (V10) increases the relative G-actin intrinsic fluorescence intensity by approximately 80% whereas for a 10-fold concentration of monomeric vanadate (V1) no effects were observed. Upon decavanadate titration, it was observed a linear increase in G-actin hydrophobic surface (2.6-fold), while no changes were detected for V1 (0-200 μM). Taken together, three major ideas arise: i) ATP prevents decavanadate-induced G-actin cysteine oxidation and vanadate reduction; ii) decavanadate promotes actin conformational changes resulting on its inactivation, iii) decavanadate has an effect on actin ATP binding site. Once it is demonstrated that actin is a new potential target for decavanadate, being the ATP binding site a suitable site for decavanadate binding, it is proposed that some of the biological effects of vanadate can be, at least in part, explained by decavanadate interactions with actin.  相似文献   

9.
Electric field pulses, capacitively applied to tissue cultures of embryonic bone cells, were shown to induce changes in the state of cellular actin. Three actin states could be defined by DNAase I inhibition. A rapidly (20-30 s) inhibiting fraction, attributed to monomeric G-actin, amounts to 55% of total actin in nonstimulated cells. An additional fraction of 8% required approx. 20 min to reach full inhibition and was tentatively defined as polymeric 'F'-actin. The remaining 37% could be detected only after treatment of the cells with 0.75 M guanidine hydrochloride, which dissociates actin from all its protein interactions. This fraction, N-actin (network actin) is believed to represent F-actin integrated into some supramolecular structure, where it is not accessible to DNAase I. Upon short electric stimulation the distribution changed to 40% G-actin, 12% F-actin and 48% N-actin. 3-Isobutyl-1-methylxanthine (IBMX; an inhibitor of cAMP phosphodiesterase), depletion of extracellular calcium, and calmodulin inhibitors abolished this field effect.  相似文献   

10.
Circular dichroic spectroscopy has been used to study the effect of pH, guanidinium hydrochloride concentration and temperature on the conformation of the fungal subtilisin-like proteinase K and the bacterial DY. The ellipticity of the bands in the far ultraviolet region remains almost unchanged in the pH range 3.0-11.0 (PMS-proteinase K) and 5.0-10.0 (PMS-subtilisin DY). The same ranges of pH stability were determined from the pH dependence of the near ultraviolet dichroic spectra. Hence the changes in the tertiary and secondary structure occur in parallel. Proteinase K is considerably more stable at acidic and somewhat more stable at alkaline pH than subtilisin DY. At neutral pH proteinase K is more resistant to denaturation by guanidinium hydrochloride than is subtilisin DY. The midpoints of the denaturation curves were 6.2 M and 3.2 M guanidinium, respectively. The thermal unfolding of proteinase K occurred at a higher temperature than for subtilisin DY, the transition midpoints being 65 degrees and 48 degrees, respectively. Thus proteinase K is overall a much more robust molecule than subtilisin DY, showing greater resistance to all three forms of denaturation. The differences in the stability of the two proteinases can be partly explained by differences in their calcium binding sites.  相似文献   

11.
We have characterized the interaction of bovine pancreatic deoxyribonuclease I (DNase I) with the filamentous (F-)actin of red cell membrane skeletons stabilized with phalloidin. The hydrolysis of [3H]DNA was used to assay DNase I. We found that DNase I bound to a homogenous class of approximately equal to 2.4 X 10(4) sites/skeleton with an association rate constant of approximately 1 X 10(6) M-1 S-1 and a KD of 1.9 X 10(-9) M at 20 degrees C. Phalloidin lowered the dissociation constant by approximately 1 order of magnitude. The DNase I which sedimented with the skeletons was catalytically inactive but could be reactivated by dissociation from the actin. Actin and DNA bound to DNase I in a mutually exclusive fashion without formation of a ternary complex. Phalloidin-treated red cell F-actin resembled rabbit muscle G-actin in all respects tested. Since the DNase I binding capacity of the skeletons corresponded to the number of actin protofilaments previously estimated by other methods, it seemed likely that the enzyme binding site was confined to one end of the filament. We confirmed this premise by showing that elongating the red cell filaments with rabbit muscle actin monomers did not appreciably add to their capacity to bind or inhibit DNase I. Saturation of skeletons with cytochalasin D or gelsolin, avid ligands for the barbed end of actin filaments, did not reduce their binding of DNase I. Furthermore, neither cytochalasin D nor DNase I alone blocked all of the sites for addition of monomeric pyrene-labeled rabbit muscle G-actin to phalloidin-treated skeletons; however, a combination of the two agents did so. In the presence of phalloidin, the polymerization of 300 nM pyrenyl actin on nuclei constructed from 5 nM gelsolin and 25 nM rabbit muscle G-actin was completely inhibited by 35 nM DNase I but not by 35 nM cytochalasin D. We conclude that DNase I associates uniquely with and caps the pointed (slow-growing or negative) end of F-actin. These results imply that the amino-terminal, DNase I-binding domain of the actin protomer is oriented toward the pointed end and is buried along the length of the actin filament.  相似文献   

12.
《The Journal of cell biology》1984,98(3):1102-1110
The directional polymerization of actin on the erythrocyte membrane has been examined at various concentrations of G-actin by thin-section electron microscopy. For this purpose, a new experimental system using single-layered erythrocyte membranes with the cytoplasmic surfaces freely exposed was developed. The preformed actin filaments did not bind with the cytoplasmic surface of the erythrocyte membranes. When the erythrocyte membranes were incubated at low concentrations (0.3 and 0.5 microM) of G-actin, greater than 80% of polymerized actin filaments pointed toward the membranes mainly in an end-on fashion, as judged by arrowhead formation with heavy meromyosin. At higher concentrations (2 and 4 microM) of G-actin, about half of the polymerized actin filaments were directed with arrowheads pointing toward the membranes, while the rest of the filaments showed the opposite polarity pointing away from the membranes. The majority of polymerized actin filaments formed loops at the points of attachment to the membranes. In contrast, when G-actin (2 and 4 microM) in the presence of cytochalasin B was polymerized into filaments, approximately 70% showed the polarity pointing away from the membrane mainly in an end-on fashion. To check the treadmilling phenomena, the erythrocyte membranes with bidirectionally polymerized actin filaments were further incubated with G-actin at the overall critical concentration. In this case, almost all (90%) of actin filaments showed the polarity with arrowheads pointing toward the membranes. The results obtained are discussed with special reference to the mode of association of actin filaments with the plasma membrane in general.  相似文献   

13.
Changes in the actin-myosin interface are thought to play an important role in microfilament-linked cellular movements. In this study, we compared the actin binding properties of the motor domain of Dictyostelium discoideum (M765) and rabbit skeletal muscle myosin subfragment-1 (S1). The Dictyostelium motor domain resembles S1(A2) (S1 carrying the A2 light chain) in its interaction with G-actin. Similar to S1(A2), none of the Dictyostelium motor domain constructs induced G-actin polymerization. The affinity of monomeric actin (G-actin) was 20-fold lower for M765 than for S1(A2) but increasing the number of positive charges in the loop 2 region of the D. discoideum motor domain (residues 613-623) resulted in equivalent affinities of G-actin for M765 and for S1. Proteolytic cleavage and cross-linking approaches were used to show that M765, like S1, interacts via the loop 2 region with filamentous actin (F-actin). For both types of myosin, F-actin prevents trypsin cleavage in the loop 2 region and F-actin segment 1-28 can be cross-linked to loop 2 residues by a carbodiimide-induced reaction. In contrast with the S1, loop residues 559-565 of D. discoideum myosin was not cross-linked to F-actin, probably due to the lower number of positive charges. These results confirm the importance of the loop 2 region of myosin for the interaction with both G-actin and F-actin, regardless of the source of myosin. The differences observed in the way in which M765 and S1 interact with actin may be linked to more general differences in the structure of the actomyosin interface of muscle and nonmuscle myosins.  相似文献   

14.
Actin pools have been evaluated in Drosophila melanogaster Kc 0% cells, through an actin assay based on differential inhibition of DNase I by globular (G) and filamentous (F) actin. Total actin represents about 4 % of total proteins and 54 % is G-actin. In ecdysterone treated cells (0.1 μM), the total actin content increases up to 9 % of total proteins after 3 days of treatment. Ecdysterone induces increase of G-actin as well as F-actin. Increase of both actins, detectable after only 24 hrs of treatment, is roughly parallel during the first two days of treatment. For longer hormonal treatment, actin polymerization is more important than accumulation of G-actin. Indirect immunofluorescence microscopy with antibodies to exogeneous DNase I suggests that actin is widely distributed in the whole cytoplasm before and after ecdysterone treatment. These results suggest that ecdysterone induces actin synthesis and polymerization in Drosophila melanogaster cells.  相似文献   

15.
Irreversible disassembly of the 4Fe-4S cluster in Chromatium vinosum high-potential iron protein (HiPIP) has been investigated in the presence of a low concentration of guanidinium hydrochloride. From the dependence of degradation rate on [H+], it is deduced that at least three protons are required to trigger efficient cluster degradation. Under these conditions the protonated cluster shows broadened M?ssbauer signals, but delta EQ (1.1 mm/s) and delta (0.44 mm/s) are similar to the native form. Collapse of the protonated transition state complex, revealed by rapid-quench M?ssbauer experiments, occurs with a measured rate constant kobs approximately 0.72 +/- 0.35 s-1 that is consistent with results from time-resolved electronic absorption and fluorescence (kobs approximately 0.4 +/- 0.1 s-1) and EPR (kobs approximately 0.62 +/- 0.18 s-1) measurements. Apparently, guanidinium hydrochloride serves to perturb the tertiary structure of the protein, facilitating protonation of the cluster, but not degradation per se. Release of iron ions occurs even more slowly with kobs approximately 0.07 +/- 0.02 s-1, as determined by the appearance of the g = 4.3 EPR signal. Proton-mediated cluster degradation is sensitive to the oxidation state of the cluster, with the oxidized state showing a two-fold slower rate in acidic solutions as a result of increased electrostatic repulsion with the cluster. Consistent results are obtained from absorption, fluorescence, M?ssbauer and EPR measurements.  相似文献   

16.
Although the number of papers about "vanadium" has doubled in the last decade, the studies about "vanadium and actin" are scarce. In the present review, the effects of vanadyl, vanadate and decavanadate on actin structure and function are compared. Decavanadate (51)V NMR signals, at -516 ppm, broadened and decreased in intensity upon actin titration, whereas no effects were observed for vanadate monomers, at -560 ppm. Decavanadate is the only species inducing actin cysteine oxidation and vanadyl formation, both processes being prevented by the natural ligand of the protein, ATP. Vanadyl titration with monomeric actin (G-actin), analysed by EPR spectroscopy, reveals a 1:1 binding stoichiometry and a K(d) of 7.5 μM(-1). Both decavanadate and vanadyl inhibited G-actin polymerization into actin filaments (F-actin), with a IC(50) of 68 and 300 μM, respectively, as analysed by light scattering assays, whereas no effects were detected for vanadate up to 2 mM. However, only vanadyl (up to 200 μM) induces 100% of G-actin intrinsic fluorescence quenching, whereas decavanadate shows an opposite effect, which suggests the presence of vanadyl high affinity actin binding sites. Decavanadate increases (2.6-fold) the actin hydrophobic surface, evaluated using the ANSA probe, whereas vanadyl decreases it (15%). Both vanadium species increased the ε-ATP exchange rate (k = 6.5 × 10(-3) s(-1) and 4.47 × 10(-3) s(-1) for decavanadate and vanadyl, respectively). Finally, (1)H NMR spectra of G-actin treated with 0.1 mM decavanadate clearly indicate that major alterations occur in protein structure, which are much less visible in the presence of ATP, confirming the preventive effect of the nucleotide on the decavanadate interaction with the protein. Putting it all together, it is suggested that actin, which is involved in many cellular processes, might be a potential target not only for decavanadate but above all for vanadyl. By affecting actin structure and function, vanadium can regulate many cellular processes of great physiological significance.  相似文献   

17.
Purified skeletal muscle actins from species whose ambient pressures range from 1 to greater than 500 atm were examined for the sensitivity to hydrostatic pressure of the globular (G) to filamentous (F) self-assembly reaction. Both the equilibrium position and the kinetics of self-assembly were affected by pressure. Increased pressure shifted the self-assembly equilibrium toward the monomer (G) state and reduced the rate of F-actin assembly. For most of the actins studied, the perturbation by pressure of F-actin formation decreased with increasing measurement of pressure, indicating that F-actin has a higher compressibility than G-actin. The increase in system volume and compressibility concomitant with the assembly of F-actin can be interpreted as reflections of the major role played by hydrophobic effects in stabilizing F-actin and of the existence of "hard" binding sites, in the terminology of Torgerson et al. [Torgerson, P. M., Drickamer, H. G., & Weber, G. (1979) Biochemistry 18, 3079-3083], in the actin subunits. For actin from the deepest occurring species studied, the teleost fish Coryphaenoides armatus, which occurs to depths of approximately 5000 m (equivalent to 501 atm of pressure), there was no difference in compressibility between G-actin and F-actin; that is, the effect of increasing pressure on self-assembly was linear over the entire pressure range examined, 600 atm. The self-assembly reaction of the actin from C. armatus also differed from that of the other actins examined in that the G to F equilibrium was relatively insensitive to increased pressure; i.e., the volume change (delta V) of assembly was small.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Numerous studies have described the F-actin cytoskeleton; however, little information relevant to C-actin is available. The actin pools of bovine aortic endothelial cells were examined using in situ and in vitro conditions and fluorescent probes for G-(deoxyribonuclease I.0.3 μM) or F-actin (phalloidin, 0.2 μM). Cells in situ displayed a diffuse G-actin distribution, while F-actin was concentrated in the cell periphery and in fine stress fibers that traversed some cells. Cells of subconfluent or just confluent cultures demonstrated intense fluorescence, with many F-actin stress fibers. Postconfluent cultures resembled the condition in situ; peripheral F-actin was prominent, traversing actin stress fibers were greatly reduced and fluorescent intensity was diminished. Postconfluency had little influence on G-actin. with only an enhancement in the intensity of G-actin punctate fluorescence. When post-confluent cultures were incubated with cytochalasin D (15 min; 10--4 M), F-actin networks were disrupted and actin punctate and diffuse fluorescence increased. G-actin fluorescence was not altered by the incubation. Although its unstructured nature may account for the minor changes observed, the stability of the G-actin pool in the presence of notable F-actin modulations suggested that filamentous actin was the key constituent involved in these actin cytoskeletal alterations. A separate finding illustrated that the concomitant use of actin probes with image enhancement and fluorescent microscopy could reveal simultaneously the G- and F-actin pools within the same cell.  相似文献   

19.
The quaternary structure of streptavidin in urea   总被引:2,自引:0,他引:2  
We report on the interactions of urea and guanidinium salts with streptavidin. Gel filtration chromatography in 0, 4, 6, and 7 M urea indicates that the streptavidin tetramer remains intact in urea. Biotin alters the electrophoretic mobility of streptavidin whether or not 6 M urea is present. The intrinsic fluorescence of streptavidin is increased and blue-shifted in 6 M urea. The fluorescence changes indicate the absence of unfolding. A conformational response to urea is possible, but much of the fluorescence change is due to urea binding as a weak biotin analog (Ka approximately 1.3 M-1). The resistance to structural perturbation by urea reflects the structural stability of streptavidin's anti-parallel beta-barrel motif. Unfolding is sluggish in 6 M guanidinium hydrochloride (half-time, approximately 50 days). After guanidinium thiocyanate unfolding, streptavidin can be refolded, but the unfolding and refolding transitions are centered at different concentrations of perturbant. Slow unfolding, with a 15th power dependence on guanidinium thiocyanate concentration, may be partially responsible for the noncoincidence of the unfolding and refolding processes. Nonequilibrium behavior is also seen in 6 M urea, as native streptavidin does not unfold and guanidinium thiocyanate unfolded streptavidin does not refold. Refolding does occur at lower concentrations of urea. Guanidinium thiocyanate only slowly unfolds the biotin-streptavidin complex. In the presence of biotin, unfolded streptavidin does not refold in 6 M guanidinium thiocyanate or in 6 M urea.  相似文献   

20.
We varied the molar ratio of added lanthanide ion to skeletal muscle actin (M3+/A) and observed their effects on the change in reduced viscosity (Nred) in the presence of polymerizing quantities of salt (0.1 M KC1). Once the concentration of the lanthanide ion exceeds the concentration of the nucleotide present (0.2 mM ATP), we noted that with M3+/A ratios up to 4: (a) there was a sharp peak in the observed Nred above the level achieved by control F-actin; (b) the magnitude of (a) was shown to be a function of the initial G-actin concentration. With an M3+/A ratio of greater than 4 we observed: (i) a sharp fall in the observed Nred; (ii) the formation of an insoluble aggregate of actin; (iii) the formation of (ii) was completely reversed by removal of the M3+; (iv) a complete inhibition of the ATP hydrolysis which always accompanies the G- to F-actin transition; (v) the number of mol of M3+ required to completely inhibit the rise in Nred (above the viscosity of G-actin) was a function of the ionic radii of the 11 lanthanide ions tested; and (vi) the effects described in (i) were not mimicked when the initial protein was in the F form. In the absence of added KCI, divalent cations (e.g. Mg2+) polymerize G-actin but this effect is not mimicked by the addition of the lanthanide ions. However, under these conditions the lanthanide ions cause the formation of an insoluble aggregate of actin. We conclude that with greater than 4 mol of lanthanide ions, G-actin aggregates in a form which contains little or no F-actin and that the lanthanide ion-induced aggregates are therefore different from the Mg2+-induced F-actin paracrystals.  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