首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Gelsolin participates in the reorganization of the actin cytoskeleton that is required during such phenomena as cell movement, cytokinesis, and apoptosis. It consists of six structurally similar domains, G1-G6, which are arranged at resting intracellular levels of calcium ion so as to obscure the three actin-binding surfaces. Elevation of Ca(2+) concentrations releases latches within the constrained structure and produces large shifts in the relative positioning of the domains, permitting gelsolin to bind to and sever actin filaments. How Ca(2+) is able to activate gelsolin has been a major question concerning the function of this protein. We present the improved structure of the C-terminal half of gelsolin bound to monomeric actin at 3.0 A resolution. Two classes of Ca(2+)-binding site are evident on gelsolin: type 1 sites share coordination of Ca(2+) with actin, while type 2 sites are wholly contained within gelsolin. This structure of the complex reveals the locations of two novel metal ion-binding sites in domains G5 and G6, respectively. We identify both as type 2 sites. The absolute conservation of the type 2 calcium-ligating residues across the six domains of gelsolin suggests that this site exists in each of the domains. In total, gelsolin has the potential to bind eight calcium ions, two type 1 and six type 2. The function of the type 2 sites is to facilitate structural rearrangements within gelsolin as part of the activation and actin-binding and severing processes. We propose the novel type 2 site in G6 to be the critical site that initiates overall activation of gelsolin by releasing the tail latch that locks calcium-free gelsolin in a conformation unable to bind actin.  相似文献   

2.
Gelsolin is an actin-binding protein that is regulated by the occupancy of multiple calcium-binding sites. We have studied calcium induced conformational changes in the G1-2 and G1-3 sub-domains, and report the binding affinities for the three type II sites. A new probe for G3 has been produced and a K(d) of 5 microM has been measured for calcium in the context of G1-3. The two halves of gelsolin, G1-3 and G4-6 bind weakly with or without calcium, suggesting that once separated by apoptotic proteolysis, G1-3 and G4-6 remain apart allowing G1-3 to sever actin in a calcium free manner.  相似文献   

3.
Gelsolin is an abundant calcium dependent actin filament severing and capping protein. In the absence of calcium the molecule is compact but in the presence of calcium, as its six similar domains alter their relative position, a generally more open configuration is adopted to reveal the three actin binding sites. It is generally held that a 'helical-latch' at the C-terminus of gelsolin's domain 6 (G6), binds domain 2 (G2) to keep gelsolin in the calcium-free compact state, and that the crutial calcium binding site(s) reside in the C-terminal half of gelsolin perhaps involving the C-terminal helix itself has to be bound to release this latch. Here we provide evidence for a calcium dependent conformational change within G2 (Kd = approximately 15 micro m). We also report a calcium dependent binding site for the C-terminus (G4-6) within G2 and delimit this further to a specific region formed by residues 203-225 and 159-193. It is known that the activation of gelsolin involves multiple calcium binding events (around 6) the first of which (in G6) may release the latch. We propose that the calcium-dependent conformational change in G2 may be a subsequent step that is necessary for the dissociation of G2 from G4-6, and that this movement occurs in sympathy with calcium induced conformational changes within G6 by the physical coupling of the two calcium binding sites within G2 and G6. Additional calcium binding in other domains then result in the complete opening and activation of the gelsolin molecule.  相似文献   

4.
Gelsolin and calponin are well characterized actin-binding proteins that form a tight gelsolin:calponin complex (GCC). We show here that the GCC is formed through two distinct interfaces. One of these is formed between 144-182 of calponin and 25-150 of gelsolin (G1). The second is a calcium-sensitive site centred on calponin's CH domain, and the C-terminal half of gelsolin (G4-6). The behaviour of this second interface is dependent on the presence of calcium and so it is possible that potential GCC-binding partners may be selected by calcium availability. Actin is one such GCC-binding partner and we show that a larger complex is formed with monomeric actin in calcium. The stoichiometry of this complex is determined to be 1 gelsolin/1 calponin/2 G-actins (GCA(2)). Both actin monomers bind the GCC through gelsolin. Both calponin and gelsolin are reported to play signaling roles in addition to their better-characterized actin-binding properties and it is possible that the GCC regulates both of these functions.  相似文献   

5.
Gelsolin is a calcium and pH-sensitive modulator of actin filament length. Here, we use X-ray crystallography to examine the extraction and exchange of calcium ions from their binding sites in different crystalline forms of the activated N and C-terminal halves of gelsolin, G1-G3 and G4-G6, respectively. We demonstrate that the combination of calcium and low pH activating conditions do not induce conformational changes in G4-G6 beyond those elicited by calcium alone. EGTA is able to remove calcium ions bound to the type I and type II metal ion-binding sites in G4-G6. Constrained by crystal contacts and stabilized by interdomain interaction surfaces, the gross structure of calcium-depleted G4-G6 remains that of the activated form. However, high-resolution details of changes in the ion-binding sites may represent the initial steps toward restoration of the arrangement of domains found in the calcium-free inactive form of gelsolin in solution. Furthermore, bathing crystals with the trivalent calcium ion mimic, Tb3+, results in anomalous scattering data that permit unequivocal localization of terbium ions in each of the proposed type I and type II ion-binding sites of both halves of gelsolin. In contrast to predictions based on solution studies, we find that no calcium ion is immune to exchange.  相似文献   

6.
Structural analysis of gelsolin domains 4-6 demonstrates that the two highest-affinity calcium ions that activate the molecule are in domains 5 and 6, one in each. An additional calcium site in domain 4 depends on subsequent actin binding and is seen only in the complex. The uncomplexed structure is primed to bind actin. Since the disposition of the three domains is similar in different crystal environments, either free or in complex with actin, the conformation in calcium is intrinsic to active gelsolin itself. Thus the actin-free structure shows that the structure with an actin monomer is a good model for an actin filament cap. The last 13 residues of domain 6 have been proposed to be a calcium-activated latch that, in the inhibited form only, links two halves of gelsolin. Comparison with the active structure shows that loosening of the latch contributes but is not central to activation. Calcium binding in domain 6 invokes a cascade of swapped ion-pairs. A basic residue swaps acidic binding partners to stabilise a straightened form of a helix that is kinked in inhibited gelsolin. The other end of the helix is connected by a loop to an edge beta-strand. In active gelsolin, an acidic residue in this helix breaks with its loop partner to form a new intrahelical ion-pairing, resulting in the breakage of the continuous sheet between domains 4 and 6, which is central to the inhibited conformation. A structural alignment of domain sequences provides a rationale to understand why the two calcium sites found here have the highest affinity amongst the five different candidate sites found in other gelsolin structures.  相似文献   

7.
Gelsolin regulates the dynamic assembly and disassembly of the actin-based cytoskeleton in non-muscle cells and clears the circulation of filaments released following cell death. Gelsolin is a six-domain (G1-G6) protein activated by calcium via a multi-step process that involves unfolding from a compact form to a more open form in which the three actin-binding sites (on the G1, G2, and G4 subdomains) become exposed. To follow the global structural changes that accompany calcium activation of gelsolin, small-angle x-ray scattering (SAXS) data were collected for full-length human plasma gelsolin at nanomolar to millimolar concentrations of free Ca2+. Analysis of these data showed that, upon increasing free Ca2+ levels, the radius of gyration (Rg) increased nearly 12 A, from 31.1+/-0.3 to 43+/-2 A, and the maximum linear dimension (Dmax) of the gelsolin molecule increased 55 A, from 100 to 155A. Structural reconstruction of gelsolin from these data provided a striking visual tracking of the gradual Ca2+-induced opening of the gelsolin molecule and highlighted the critical role played by the flexible linkers between homologous domains. The tightly packed architecture of calcium-free gelsolin, seen from both SAXS and x-ray crystallographic models, is already partially opened up in as low as 0.5 nM Ca2+. Our data confirm that, although the molecule springs open from 0 to 1 microM free Ca2+, even higher calcium concentrations help to stabilize a more open structure, with increases in Rg and Dmax of approximately 2 and approximately 15 A, respectively. At these higher calcium levels, the SAXS-based models provide a molecular shape that is compatible with that of the crystal structures solved for Ca2+/gelsolin C-terminal and N-terminal halves+/-monomeric G-actin. Placement of these crystal structures within the boundaries of the SAXS-based model suggests a movement of the G1/G2 subunits that would be required upon binding to actin.  相似文献   

8.
Gelsolin is a key actin cytoskeleton-modulating protein primarily regulated by calcium and phosphoinositides. In addition, low pH has also been suggested to activate gelsolin in the absence of Ca2+ ions, although no structural insight on this pathway is available except for a reported decrement in its diffusion coefficient at low pH. We also observed ∼1.6-fold decrease in the molecular mobility of recombinant gelsolin when buffer pH was lowered from 9 to 5. Analysis of the small angle x-ray scattering data collected over the same pH range indicated that the radius of gyration and maximum linear dimension of gelsolin molecules increased from 30.3 to 34.1 Å and from 100 to 125 Å, respectively. Models generated for each dataset indicated that similar to the Ca2+-induced process, low pH also promotes unwinding of this six-domain protein but only partially. It appeared that pH is able to induce extension of the G1 domain from the rest of the five domains, whereas the Ca2+-sensitive latch between G2 and G6 domains remains closed. Interestingly, increasing the free Ca2+ level to merely ∼40 nm, the partially open pH 5 shape “sprung open” to a shape seen earlier for this protein at pH 8 and 1 mm free Ca2+. Also, pH alone could induce a shape where the g3-g4 linker of gelsolin was open when we truncated the C-tail latch from this protein. Our results provide insight into how under physiological conditions, a drop in pH can fully activate the F-actin-severing shape of gelsolin with micromolar levels of Ca2+ available.  相似文献   

9.
Gelsolin is one of the best known actin-binding proteins with several distinct activities regulated by calcium. Using a kinase fraction isolated from mitotic HeLa cells, we found that the plasma form of gelsolin can be phosphorylated at a site located within the NH2-terminus region which does not exist in the cytoplasmic form. After this phosphorylation, gelsolin no longer requires Ca2+ for activity; it severs and subsequently caps actin filaments, and nucleates filament formation in Ca2+-free solution. These findings may clarify the mechanism of gelsolin regulation by Ca2+, and indicate that changes in electrical interactions between the NH2- and COOH-terminal ends are important for this regulation. Moreover, since only a single site is phosphorylated, and since the phosphorylated region does not contribute to this protein's own activity, the results suggest that a single chemical charge modification at a site away from the protein's core structure, such as this phosphorylation site, is sufficient to alter the protein's function.  相似文献   

10.
Villin and gelsolin consist of six homologous domains of the gelsolin/cofilin fold (V1–V6 and G1–G6, respectively). Villin differs from gelsolin in possessing at its C terminus an unrelated seventh domain, the villin headpiece. Here, we present the crystal structure of villin domain V6 in an environment in which intact villin would be inactive, in the absence of bound Ca2+ or phosphorylation. The structure of V6 more closely resembles that of the activated form of G6, which contains one bound Ca2+, rather than that of the calcium ion-free form of G6 within intact inactive gelsolin. Strikingly apparent is that the long helix in V6 is straight, as found in the activated form of G6, as opposed to the kinked version in inactive gelsolin. Molecular dynamics calculations suggest that the preferable conformation for this helix in the isolated G6 domain is also straight in the absence of Ca2+ and other gelsolin domains. However, the G6 helix bends in intact calcium ion-free gelsolin to allow interaction with G2 and G4. We suggest that a similar situation exists in villin. Within the intact protein, a bent V6 helix, when triggered by Ca2+, straightens and helps push apart adjacent domains to expose actin-binding sites within the protein. The sixth domain in this superfamily of proteins serves as a keystone that locks together a compact ensemble of domains in an inactive state. Perturbing the keystone initiates reorganization of the structure to reveal previously buried actin-binding sites.Actin is crucial to such processes as cell movement, cell division, and apoptosis, which are regulated by numerous actin-binding proteins, including gelsolin, Arp2/3, and profilin (for review, see Ref. 1). Gelsolin, the most potent actin filament-severing protein known, can bind to, sever, cap, and nucleate actin filaments in a calcium-, pH-, ATP-, and phospholipid-dependent manner (for review, see Ref. 2). Villin, found in microvilli of absorptive epithelium, is a second member of the gelsolin family of actin-binding proteins. In addition to standard gelsolin-type activities, villin is able to bundle actin filaments and is subject to regulation by tyrosine phosphorylation as well as by Ca2+ and phosphatidylinositol 4,5-bisphosphate (for review, see Ref. 3). Many comparisons have been made between gelsolin and villin. The two share 50% amino acid sequence identity and show similar proteolytic cleavage patterns (4). Both contain six similarly folded domains, but villin possesses a seventh domain at its C terminus, the headpiece (HP)2 domain, which folds into a compact structure that introduces a second F-actin-binding site into the protein. Recent studies indicate that villin uses the HP F-actin-binding sites to achieve bundling (5). In an environment devoid of free Ca2+, gelsolin and villin assume inactive conformations. After binding Ca2+, both undergo conformational rearrangements that expose their binding sites for F-actin. In villin, this includes revealing the HP actin-binding site through a “hinge mechanism” (6).Biochemical and structural studies have revealed eight Ca2+-binding sites of two types in gelsolin (for review, see Ref. 7). Each of the six domains contains a complete and evolutionarily conserved site, termed type 2, whereas G1 and G4 provide partial Ca2+ coordination at interfaces with actin through sites termed type 1. Sequential mutagenesis of these sites in villin has identified six functional Ca2+-binding sites (8): two major sites, one each of type 1 and type 2, in V1, plus four type 2 sites in V2–V6. The type 1 site in V1 regulates F-actin-capping and F-actin-severing activities, whereas the lower affinity type 2 site in V1 only affects severing (9). The other four sites are involved in stabilizing villin conformation, but they do not directly influence actin-severing activity. NMR studies of a fragment of villin that consists of V6 and the HP domain have implicated V6 residues Asn647, Asp648, and Glu670 in binding Ca2+ (10). These experiments also revealed the first 80 residues of V6 to undergo significant conformational change as a result of Ca2+ binding.Nanomolar to micromolar concentrations of free Ca2+ govern the actin-binding activities of gelsolin. In contrast, micromolar and millimolar concentrations of calcium ions are required for villin to exhibit capping and severing, respectively. However, after tyrosine phosphorylation, villin can sever actin filaments even at nanomolar Ca2+ concentrations (11). Furthermore, although the actin-severing ability of the N-terminal half of villin is calcium-dependent, that by the N-terminal half of gelsolin is not. In contrast, the binding of G-actin of the C-terminal half of both villin and gelsolin requires Ca2+. Creation of hybrid proteins demonstrated that the domains of villin and gelsolin are not interchangeable (12).Abundant x-ray crystallographic structural information exists for gelsolin, including the calcium ion-free (Ca2+-free), inactive structure of the intact protein (13), the activated N- and C-terminal halves, each in a bimolecular complex with actin (7, 14), and the activated C-terminal half on its own (15, 16). Structural data for intact villin are unavailable and are limited to fragment V1 (17), solved using NMR methods, and the HP domain, solved by NMR and x-ray crystallography (18, 19). NMR experiments also indicate that HP is connected to V6 by a 40-residue disordered linker. As a result, HP has been proposed to bind actin independently of the remainder of the protein (10).In this report, we present the structure of Ca2+-free, isolated villin V6, which exhibits a typical gelsolin domain fold. The long helix in V6 in this structure is straight, unlike the corresponding helix in G6 of intact Ca2+-free gelsolin, which is bent, and only straightens on calcium activation of the intact protein. Hence, V6 appears to be in an active conformation in the absence of Ca2+. Molecular dynamics simulations indicate that the preferred state of the long helix is also straight for isolated G6 in the absence of Ca2+. Furthermore, they suggest a bistable mechanism of helix conformational change regulated by the presence of the remaining domains, by calcium ions, and by other interactants. We therefore propose a mechanism for the gelsolin family proteins whereby Ca2+ triggers the straightening of the domain 6 helix in the native conformation of the inactive proteins to propagate more widespread conformational changes.  相似文献   

11.
Gelsolin is a Ca2+-regulated actin-binding protein that can sever, cap, and nucleate growth from the pointed ends of actin filaments. In this study we have measured the binding of the amino-terminal half of gelsolin, G1-3, to pyrene-labeled F-actin as a function of Ca2+ concentration. The rate of binding is shown to be dependent on micromolar concentrations of Ca2+. Independent experiments demonstrate that conformational changes in G1-3 are induced by micromolar concentrations of Ca2+. Titrations of pyrene-F-actin with G1-3 and gelsolin show that the quenching of pyrene fluorescence is identical in extent and stoichiometry for both G1-3 and gelsolin. In contrast, severing of F-actin by G1-3 is found to be much less efficient than is severing by gelsolin. In experiments in which F-actin severing is quantitatively measured, the filament number is found to be proportional to the 1.35 power of the G1-3 concentration. This deviation from linearity may be explained by cooperativity; the binding of two G1-3 molecules in close proximity may lead to cooperative severing of the polymer, thus increasing the severing efficiency. This model is supported by experiments that show that the efficiency of G1-3 severing of F-actin increases with increasing G1-3:F-actin ratios. Extrapolating from these results, we conclude that G4-6, the carboxyl-terminal half of gelsolin, has an active role in the severing of F-actin by intact gelsolin. Whereas F-actin severing by G1-3 is enhanced by cooperative binding of two separate G1-3 molecules, cooperativity is inherent to intact gelsolin because the cooperative partners are covalently linked.  相似文献   

12.
pH dependence of hydrogen exchange from backbone peptide amides in apamin   总被引:1,自引:0,他引:1  
C E Dempsey 《Biochemistry》1986,25(13):3904-3911
The kinetics of hydrogen exchange of the 11 most protected backbone amides of bee venom apamin have been measured between pH 1 and pH 8.5 by using time-resolved and saturation-transfer NMR spectroscopy. The five amides most protected from base-catalyzed exchange, those of residues 5 and 12-15, show highly correlated exchange behavior in the base-catalyzed regime. It is proposed that the intramolecular hydrogen bonds stabilizing these amides define a stable cooperative unit of secondary structure in apamin (a C-terminal helix and an N-terminal beta-turn). This conformational unit is further stabilized (by 5-6 kJ mol-1) on titration of the Glu-7 side-chain carboxyl group. The relative contributions of specific intramolecular interactions to this conformational stabilization are estimated. The pHminima in the pH-dependent single amide exchange curves are compared with values predicted by correcting for sequence-dependent contributions to amide exchange rates [Molday, R. S., Englander, S. W., & Kallen, R. G. (1972) Biochemistry 11, 150-158]. The lack of correlation suggests that the "open" conformers from which amide exchange occurs are nonrandom. This conclusion is dependent on the assumption that acid-catalyzed exchange occurs via N-protonation so that residual conformational effects on exchange rates in the open conformers will affect acid- and base-catalyzed rates in approximately equal and opposite ways. A strong correlation between the measured pHminima and the amide proton chemical shifts is observed, however, and this may be most easily accommodated if acid-catalyzed exchange occurs by the imidic acid mechanism (via amide O-protonation).  相似文献   

13.
Because of its ability to rapidly depolymerize F-actin, plasma gelsolin has emerged as a therapeutic molecule in different disease conditions. High amounts of exogenous gelsolin are, however, required to treat animal models of different diseases. Knowing that the F-actin depolymerizing property of gelsolin resides in its N terminus, we made several truncated versions of plasma gelsolin. The smaller versions, particularly the one composed of the first 28–161 residues, depolymerized the F-actin much faster than the native gelsolin and other truncates at the same molar ratios. Although G1-G3 loses its dependence on Ca2+ or low pH for the actin depolymerization function, interestingly, G1-G2 and its smaller versions were found to regain this requirement. Small angle x-ray scattering-based shape reconstructions revealed that G1-G3 adopts an open shape in both the presence and the absence of Ca2+ as well as low pH, whereas G1-G2 and residues 28–161 prefer collapsed states in Ca2+-free conditions at pH 8. The mutations in the g2-g3 linker resulted in the calcium sensitivity of the mutant G1-G3 for F-actin depolymerization activity, although the F-actin-binding sites remained exposed in the mutant G1-G3 as well as in the smaller truncates even in the Ca2+-free conditions at pH 8. Furthermore, unlike wild type G1-G3, calcium-sensitive mutants of G1-G3 acquired closed shapes in the absence of free calcium, implying a role of g2-g3 linker in determining the open F-actin depolymerizing-competent shape of G1-G3 in this condition. We demonstrate that the mobility of the G1 domain, essential for F-actin depolymerization, is indirectly regulated by the gelsolin-like sequence of g2-g3 linker.  相似文献   

14.
To investigate the pH dependence of the conformational stability of barnase, urea denaturation curves were determined over the pH range 2-10. The maximum conformational stability of barnase is 9 kcal mol-1 and occurs between pH 5 and 6. The dependence of delta G on urea concentration increases from 1850 cal mol-1 M-1 at high pH to about 3000 cal mol-1 M-1 near pH 3. This suggests that the unfolded conformations of barnase become more accessible to urea as the net charge on the molecule increases. Previous studies suggested that in 8 M urea barnase unfolds more completely than ribonuclease T1, even with the disulfide bonds broken [Pace, C.N., Laurents, D. V., & Thomson, J.A. (1990) Biochemistry 29, 2564-2572]. In support of this, solvent perturbation difference spectroscopy showed that in 8 M urea the Trp and Tyr residues in barnase are more accessible to perturbation by dimethyl sulfoxide than in ribonuclease T1 with the disulfide bonds broken.  相似文献   

15.
Gelsolin is an actin filament severing protein composed of six similar structured domains that differ with respect to actin, calcium and polyphospho-inositide binding. Previous work has established that gelsolin binds tropomyosin [Koepf, E.K. and Burtnick, L.D. (1992) FEBS Lett. 309, 56-58]. We have produced various specific gelsolin domains in Escherichia coli in order to establish which of the six domains binds tropomyosin. Gelsolin domains 1-3 (G1-3), G1-2 and G2 all bind tropomyosin in a pH and calcium insensitive manner whereas binding of G4-6 to tropomyosin was barely detectable under the conditions tested. We conclude that gelsolin binds tropomyosin via domain 2 (G2).  相似文献   

16.
A marked pH-dependent interaction with F-actin is an important property of typical members of the actin depolymerizing factor (ADF)/cofilin family of abundant actin-binding proteins. ADF/cofilins tend to bind to F-actin with a ratio of 1 : 1 at pH values around 6.5, and to G-actin at pH 8.0. We have investigated the mechanism for the pH-sensitivity. We found no evidence for pH-dependent changes in the structure of cofilin itself, nor for the interaction of cofilin with G-actin. None of the actin-derived, cofilin-binding peptides that we had previously identified [Renoult, C., Ternent, D., Maciver, S.K., Fattoum, A., Astier, C., Benyamin, Y. & Roustan, C. (1999) J. Biol. Chem. 274, 28893-28899] bound cofilin in a pH-sensitive manner. However, we have detected a conformational change in region 75-105 in the actin subdomain 1 by the use of a peptide-directed antibody. A pH-dependent conformational change has also been detected spectroscopically in a similar peptide (84-103) on binding to cofilin. These results are consistent with a model in which pH-dependent motion of subdomain 1 relative to subdomain 2 (through region 75-105) of actin reveals a second cofilin binding site on actin (centered around region 112-125) that allows ADF/cofilin association with the actin filament. This motion requires salt in addition to low pH.  相似文献   

17.
《The Journal of cell biology》1985,101(4):1236-1244
Platelet gelsolin (G), a 90,000-mol-wt protein, binds tightly to actin (A) and calcium at low ionic strength to form a 1:2:2 complex, GA2Ca2 (Bryan, J., and M. Kurth, 1984, J. Biol. Chem. 259:7480-7487). Chromatography of actin and gelsolin mixtures in EGTA-containing solutions isolates a stable binary complex, GA1Ca1 (Kurth, M., and J. Bryan, 1984, J. Biol. Chem. 259:7473-7479). The effects of platelet gelsolin and the binary gelsolin-actin complex on the depolymerization kinetics of rabbit skeletal muscle actin were studied by diluting pyrenyl F-actin into gelsolin or complex-containing buffers; a decrease in fluorescence represents disassembly of filaments. Dilution of F- actin to below the critical concentration required for filament assembly gave a biphasic depolymerization curve with both fast and slow components. Dilution into buffers containing gelsolin, as GCa2, increased the rate of depolymerization and gave a first order decay. The rate of decrease in fluorescence was found to be gelsolin concentration dependent. Electron microscopy of samples taken shortly after dilution into GCa2 showed a marked reduction in filament length consistent with filament severing and an increase in the number of ends. Conversely, occupancy of the EGTA-stable actin-binding site by an actin monomer eliminated the severing activity. Dilution of F-actin into the gelsolin-actin complex, either as GA1Ca1 or GA1Ca2, resulted in a decrease in the rate of depolymerization that was consistent with filament end capping. This result indicates that the EGTA-stable binding site is required and must be unoccupied for filament severing to occur. The effectiveness of gelsolin, GCa2, in causing filament depolymerization was dependent upon the ionic conditions: in KCI, actin filaments appeared to be more stable and less susceptible to gelsolin, whereas in Mg2+, actin filaments were more easily fragmented. Finally, a comparison of the number of kinetically active ends generated when filaments were diluted into gelsolin versus the number formed when gelsolin can function as a nucleation site suggests that gelsolin may sever more than once. The data are consistent with a mechanism where gelsolin, with both actin-binding sites unoccupied, can sever but not cap F-actin. Occupancy of the EGTA-stable binding site yields a gelsolin-actin complex that can no longer sever filaments, but can cap filament ends.  相似文献   

18.
Gelsolin requires activation to carry out its severing and capping activities on F-actin. Here, we present the structure of the isolated C-terminal half of gelsolin (G4-G6) at 2.0 A resolution in the presence of Ca(2+) ions. This structure completes a triptych of the states of activation of G4-G6 that illuminates its role in the function of gelsolin. Activated G4-G6 displays an open conformation, with the actin-binding site on G4 fully exposed and all three type-2 Ca(2+) sites occupied. Neither actin nor the type-l Ca(2+), which normally is sandwiched between actin and G4, is required to achieve this conformation.  相似文献   

19.
Unfertilized eggs of the medaka fish (Oryzias latipes) were injected with pH-buffered calcium buffers. Medaka egg activation is accompanied by a transient increase in cytoplasmic free calcium (Gilkey, J. C., L. F. Jaffe, E. B. Ridgway, and G. T. Reynolds, 1978, J. Cell Biol., 76:448-466). The calcium buffer injections demonstrated that (a) the threshold free calcium required to elicit the calcium transient and activate the egg is between 1.7 and 5.1 microM at pH 7.0, well below the 30 microM reached during the transient, and (b) buffers which hold free calcium below threshold prevent activation of the buffered region in subsequently fertilized eggs. Therefore an increase in free calcium is necessary and sufficient to elicit the calcium transient, and the calcium transient is necessary to activate the egg. Further, these results are additional proof that the calcium transient is initiated and propagated through the cytoplasm by a mechanism of calcium- stimulated calcium release. Finally, a normal calcium transient must propagate through the entire cytoplasm to ensure normal development. Unfertilized eggs were injected with pH buffers to produce short-term, localized changes in cytoplasmic pH. The eggs were then fertilized at various times after injection. In other experiments, unfertilized and fertilized eggs were exposed to media containing either NH4Cl or CO2 to produce longer term, global changes in cytoplasmic pH. These treatments neither activated the eggs nor interfered with the normal development of fertilized eggs, suggesting that even if a natural change in cytoplasmic pH is induced by activation, it has no role in medaka egg development. The injected pH buffers altered the rate of propagation of the calcium transient through the cytoplasm, suggesting that the threshold free calcium required to trigger calcium-stimulated calcium release might be pH dependent. The results of injection of pH-buffered calcium buffers support this conjecture: for a tenfold increase in hydrogen ion concentration, free calcium must also be raised tenfold to elicit the calcium transient.  相似文献   

20.
The photosystem-II-associated 33-kDa extrinsic manganese-stabilizing protein is found in all oxygen-evolving organisms. In this paper, we show that this protein undergoes pH-induced conformational changes in the physiological pH range. At a neutral pH of 7.2, the hydrophobic amino acid residues that are most likely located inside the beta barrel are "closed" and the protein binds neither Mn2+ nor Ca2+ ions. When the protein is transferred to a solution with a slightly acidic pH of 5.7, hydrophobic amino acid residues become exposed to the surrounding medium, enabling them to bind the fluorescent probe 8,1-ANS. At this pH-induced open state, Mn2+ and Ca2+ bind to the manganese-stabilizing protein. The pH values used in this study, 7.2 and 5.7, are typical of the pH found in the thylakoid lumen in the dark and light, respectively. A model is presented in which the manganese-stabilizing protein undergoes a pH-dependent conformational change that in turn influences its capacity to bind calcium and manganese. In this model, the proton-dependent conformational changes of the tertiary structure of the manganese-stabilizing protein are of functional relevance for the regulation of substrate (water) delivery to and product (proton) release from the water-oxidizing complex by forming a proton-sensing proton-transport pathway.  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