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.
In the presence of MgATP or MgADP the E. coli chaperonin proteins, GroEL and GroES, form a stable asymmetric complex with a stoichiometry of two GroEL7:one GroES7: seven MgADP. The distribution of the ligands between the two heptameric GroEL rings is crucial to our understanding of the mechanism of chaperonin-assisted folding, being either cis (i.e. [GroEL7.MgADP7.GroES7]-[GroEL7]) or trans (i.e. [GroEL7.MgADP7]-[GroEL7.GroES7]. On the basis of cross-linking experiments with 8-azido-ATP and the heterobifunctional reagent, N-succinimidyl 3-(2-pyridyldithio) propionate (SPDP), it was suggested that GroES and MgADP are bound to the same GroEL ring which resists proteinase K digestion [Nature 366 (1993) 228-233]. However, we find that the SPDP-promoted cross linking of GroES and GroEL occurs in the absence of Mg2+, ADP or ATP, which are required for the formation of the asymmetric complex. Cross-linking is shown to occur only when the SPDP-modified GroES is co-precipitated with GroEL by trichloracetic acid. Furthermore, there are structural grounds for questioning whether SPDP can crosslink, in a physiologically relevant manner, an amino group of GroES with any of the cysteinyl groups of GroEL.  相似文献   

2.
We have previously shown that the C-terminal sequence of GroES is required for oligomerization [Seale and Horowitz (1995), J. Biol. Chem. 270, 30268-30270]. In this report, we have generated a C-terminal deletion mutant of GroES with a significantly destabilized oligomer and have investigated its function in the chaperonin-assisted protein folding cycle. Removal of the two C-terminal residues of GroES results in a cochaperonin [GroESD(96-97)] that is monomeric at concentrations where GroES function is assessed. Using equilibrium ultracentrifugation, we measured the dissociation constant for the oligomer-monomer equilibrium to be 7.3 x 10(-34)M6. The GroESD(96-97) is fully active as a cochaperonin. This mutant is able to inhibit the ATPase activity of GroEL to levels comparable to wild-type GroES. It is also able to assist the refolding of urea-denatured rhodanese by GroEL. While GroESD(96-97) can function at levels comparable to wild-type GroES, higher concentrations of mutant are required to produce the same effect. These results support the idea that the performed GroES heptamer is not required for function, but they suggest that the oligomeric cochaperonin is most efficient.  相似文献   

3.
In human hemoglobin (Hb) the beta37 tryptophan residue (betaW37), located at the hinge region of the alpha1beta2 interface, forms many contacts with alpha subunit residues of the opposite dimer, in both the T and R quaternary structures. We have carried out equilibrium O2 binding studies on a series of recombinant Hbs that have mutations at this residue site: betaW37Y, betaW37A, betaW37G, and betaW37E. Binding isotherms measured at high concentrations of these mutants were found to be shifted toward increased affinity and decreased cooperativity from that of the normal HbA0 tetramer. Analysis of these binding isotherms indicated that amino acid substitutions at the beta37 position could both destabilize the tetrameric form of the mutants relative to their constituent dimers and also alter cooperativity of the intact tetrameric species. These alterations from wild-type function are dependent on the particular side chain substituted, with the magnitude of change increasing as Trp is substituted by Tyr, Ala, Gly, and Glu. The dimer to tetramer assembly free energy of deoxy-betaW37E, the most perturbed mutant in the series, was measured using analytical gel chromatography to be 9 kcal/tetramer less favorable than that of deoxy HbA0. Stabilizing the betaW37E tetramer by addition of IHP, or by cross-linking at the alphaK99 positions, does not restore normal O2 binding behavior. Thermodynamic parameters of all the mutants were found to correlate with their CO binding rates and with their high-resolution X-ray crystal structures (see accompanying papers: Kwiatkowski et al. (1998) Biochemistry 37, 4325-4335; Peterson & Friedman (1998) Biochemistry 37, 4346-4357; Kavanaugh et al. (1998) Biochemistry 37, 4358-4373].  相似文献   

4.
Modification of the Escherichia coli chaperonin GroEL with N-ethylmaleimide at residue Cys138 affects the structural and functional integrity of the complex. Nucleotide affinity and ATPase activity of the modified chaperonin are increased, whereas cooperativity of ATP hydrolysis and affinity for GroES are reduced. As a consequence, release and folding of substrate proteins are strongly impaired and uncoupled from ATP hydrolysis in a temperature-dependent manner. Folding of dihydrofolate reductase at 25 degrees C becomes dependent on GroES, whereas folding of typically GroES-dependent proteins is blocked completely. At 37 degrees C, GroES binding is restored to normal levels, and the modified GroEL regains its chaperone activity to some extent. These results assign a central role to the intermediate GroEL domain for transmitting conformational changes between apical and central domains, and for coupling ATP hydrolysis to productive protein release.  相似文献   

5.
Cys-scanning mutagenesis has been applied to the remaining 45 residues in lactose permease that have not been mutagenized previously (from Gln100 to Arg144 which comprise helix IV and adjoining loops). Of the 45 single-Cys mutants, 26 accumulate lactose to > 75% of the steady state observed with Cys-less permease, and 14 mutants exhibit lower but significant levels of accumulation (35-65% of Cys-less permease). Permease with Phe140-->Cys or Lys131-->Cys exhibits low activity (15-20% of Cys-less permease), while mutants Gly115-->Cys, Glu126-->Cys and Arg144-->Cys are completely unable to accumulate the dissacharide. However, Cys-less permease with Ala or Pro in place of Gly115 is highly active, and replacement of Lys131 or Phe140 with Cys in wild-type permease has a less deleterious effect on activity. In contrast, mutant Glu126-->Cys or Arg144-->Cys is inactive with respect to both uphill and downhill transport in either Cys-less or wild-type permease. Furthermore, mutants Glu126-->Ala or Gln and Arg144-->Ala or Gln are also inactive in both backgrounds, and activity is not rescued by double neutral replacements or inversion of the charged residues at these positions. Finally, a mutant with Lys in place of Arg144 accumulates lactose to about 25% of the steady state of wild-type, but at a slow rate. Replacement of Glu126 with Asp, in contrast, has relatively little effect on activity. None of the effects can be attributed to decreased expression of the mutants, as judged by immunoblot analysis. Although the activity of most of the single-Cys mutants is unaffected by N-ethylmaleimide, Cys replacement at three positions (Ala127, Val132, or Phe138) renders the permease highly sensitive to alkylation. The results indicate that the cytoplasmic loop between helices IV and V, where insertional mutagenesis has little effect on activity [McKenna, E., et al. (1992) Proc. Natl. Acad. Sci. U.S.A. 89, 11954-11958], contains residues that play an important role in permease activity and that a carboxyl group at position 126 and a positive charge at position 144 are absolutely required.  相似文献   

6.
Three transmembrane glutamic acid residues play essential roles in the metal-tetracycline/H+ antiporter Tet(K) of Staphylococcus aureus [Fujihira et al., FEBS Lett. 391 (1996) 243-246]. In the putative hydrophilic loop region of the Tet(K) and Tet(L) proteins, six acidic residues are conserved. Asp74, Asp200, Asp318 and Glu381 are located on the putative cytoplasmic side, and Asp39 and Glu345 on the putative periplasmic side. These residues were replaced by a neutral amino acid residue or a charge-conserved one. In contrast to the transmembrane glutamic acid residues, the replacement of the two glutamic acid residues (Glu345 and Glu381) did not affect the tetracycline resistance level. Out of the other four aspartic acid residues, the only essential residue is Asp318, any replacement of which resulted in complete loss of the tetracycline resistance and transport activity. Asp318 is located in cytoplasmic loop 10-11 in the putative 14-transmembrane-segment topology of Tet(K). In the case of the tetracycline exporters of Gram-negative bacteria, the only essential acidic residue in the cytoplasmic loop region is located in loop 2-3 [Yamaguchi et al., Biochemistry 31 (1992) 8344-8348]. It may be a general role for tetracycline efflux proteins that three transmembrane and one cytoplasmic acidic residues are mandatory for the tetracycline transport function.  相似文献   

7.
Although the chaperonin GroEL/GroES complex binds and hydrolyzes ATP, its structure is unlike other known ATPases. In order to better characterize its nucleotide binding sites, we have photolabeled the complex with the affinity analog 2-azido-ATP. Three residues of GroEL, Pro137, Cys138 and Thr468, are labeled by the probe. The location of these residues in the GroEL crystal structure [Braig, K., Otwinowski, Z., Hedge, R., Boisvert, D., Joachimiak, A., Horwich, A. & Sigler, P. (1994) Nature 371, 578-586: Boisvert, D. C., Wang, J., Otwinowski, Z., Horwich, A. L. & Sigler, P. B. (1996) Nat. Struct. Biol. 3, 170-177] suggests that 2-azido-ATP binds to an alternative conformer of GroEL in the presence of GroES. The labeled site appears to be located at the GroEL/GroEL subunit interface since modification of Pro137 and Cys138 is most readily explained by attack of a probe molecule bound to the adjacent GroEL subunit. Labeling of the co-chaperonin, GroES, is clearly demonstrated on gels and the covalent tethering of nucleotide allows detection of a GroES dimer in the presence of SDS. However, no stable peptide derivative of GroES could be purified for sequencing. In contrast, the GroES homolog, yeast cpn10, does give a stable derivative. The modified amino acid is identified as the conserved Pro13, which corresponds to Pro5 in Escherichia coli GroES.  相似文献   

8.
Recently a number of mutations have been found in vitro which maintain alpha 1-adrenergic receptors (ARs) in a partially activated form. We have previously identified two amino acid residue positions in the alpha 1b-adrenergic receptor (AR), Cys128 and Ala204, one in each of the third and fifth transmembrane segments, that constitutively activate the receptor when substituted for a phenylalanine or valine, respectively [Perez et al. (1996) Mol. Pharmacol. 49, 112-122; Hwa et al. (1996) J. Biol. Chem. 271, 7956-7964]. Another mutation analyzed previously, Ala293Glu, located in the third intracellular loop, also constitutively activates the receptor [Kjelsborg et al. (1992) J. Biol. Chem. 267, 1430-1433]. All three mutations displayed similar manifestations of constitutive activity such as higher binding affinity and potency for agonists as well as higher basal signal transduction as predicted by the revised ternary complex model of receptor activation. We hypothesized that the individual mutations because of their critical location alter the conformation of the transmembrane helices such that mimicry occur that partially conforms to the activated state, R*. To explore whether these potential conformations are independent, we combined these three mutations in all possible permutations. The combined triple mutation displays 700-fold higher binding affinity for (-)-epinephrine and 20-fold higher basal IP3 release than the wild-type receptor. We also observed that each mutation contributed independently and synergistically to both receptor agonist binding and activation with the combined mutations basal activity exceeding that of the fully-stimulated wild-type receptor. There was also a direct correlation between epinephrine's binding affinity and the degree of constitutive activity. Because the mutations affect different transmembrane domains, these results are consistent with a mechanism that helical movement acts in a concerted fashion in agonist-induced activation, a synergism predicted if multiple helix movement is involved in receptor activation.  相似文献   

9.
Three cold-sensitive mutants in phage P22 coat protein have been characterized to determine the effects of the amino acid substitutions that cause cold sensitivity on the folding pathway and the conformation of refolded coat protein. Here we find that the three cold-sensitive mutants which have the threonine residue at position 10 changed to isoleucine (T10I), the arginine residue at position 101 changed to cysteine (R101C), or the asparagine residue at position 414 changed to serine (N414S) were capable of folding from a denatured state into a soluble monomeric species, but in each case, the folded conformation was altered. Changes in the kinetics of folding were observed by both tryptophan and bisANS fluorescence. In contrast to the temperature-sensitive for folding coat protein mutants which can be rescued at nonpermissive temperatures in vivo by the overproduction of molecular chaperones GroEL and GroES [Gordon, C. L., Sather, S. K., Casjens, S., & King, J. (1994) J. Biol. Chem. 269, 27941-27951], the folding defects associated with the cold-sensitive amino acid substitutions were not recognized by GroEL and GroES.  相似文献   

10.
Glu126 and Arg144 in the lactose permease are indispensable for substrate binding and probably form a charge-pair [Venkatesan, P., and Kaback, H. R. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 9802-9807]. Mutants with Glu126-->Ala or Arg144-->Ala do not bind ligand or catalyze lactose accumulation, efflux, exchange, downhill lactose translocation, or lactose-induced H+ influx. In contrast, mutants with conservative mutations (Glu126-->Asp or Arg144-->Lys) exhibit drastically different phenotypes. Arg144-->Lys permease accumulates lactose slowly to low levels, but does not bind ligand or catalyze equilibrium exchange, efflux, or lactose-induced H+ influx. In contrast, Glu126-->Asp permease catalyzes lactose accumulation and lactose-induced H+ influx to wild-type levels, but at significantly lower rates. Surprisingly, however, no significant exchange or efflux activity is observed. Glu126-->Asp permease exhibits about a 6-fold increase in the Km for active transport relative to wild-type permease with a comparable Vmax. Direct binding assays using flow dialysis demonstrate that mutant Glu126-->Asp binds p-nitrophenyl-alpha,D-galactopyranoside. Indirect binding assays utilizing substrate protection against [14C]-N-ethylmaleimide labeling of single-Cys148 permease reveal an apparent Kd of 3-5 mM for lactose and 15-20 microM for beta, D-galactopyranosyl-1-thio-beta,D-galactopyranoside (TDG). The affinity of Glu126-->Asp/Cys148 permease for lactose is markedly decreased (Kd > 80 mM), while TDG affinity is altered to a much lesser extent (Kd ca. 80 microM). The results extend the conclusion that a carboxylate at position 126 and a guanidinium group at position 144 are irreplaceable for substrate binding and support the idea that Arg144 plays a major role in substrate specificity.  相似文献   

11.
When chaperonins GroEL and GroES are incubated under functional conditions in the presence of ATP (5 mM) and K+ (150 mM), GroEL-GroES complexes appear in the incubation mixture, that are either asymmetric (1:1 GroEL:GroES oligomer ratio) or symmetric (1:2 GroEL:GroES oligomer ratio). The percentage of symmetric complexes present is directly related to the [ATP]/[ADP] ratio and to the K+ concentration. Kinetic analysis shows that there is a cycle of formation and disappearance of symmetric complexes. A correlation between the presence of symmetric complexes in the incubation mixture and its rhodanese folding activity suggests some active role of these complexes in the protein folding process. Accordingly, under functional conditions, symmetric complexes are found to contain denatured rhodanese. These data suggest that binding of substrate inside the GroEL cavity takes place before the symmetric complex is formed.  相似文献   

12.
In this work we show that the GroEL (Hsp60 equivalent) chaperone protein can protected purified Escherichia coli RNA polymerase (RNAP) holoenzyme from heat inactivation better than the DnaK (Hsp70 equivalent) chaperone can. In this protection reaction, the GroES protein is not essential, but its presence reduces the amount of GroEL required. GroEL and GroES can also reactivate heat-inactivated RNAP in the presence of ATP. The mutant GroEL673 protein, with or without GroES, is incapable of reactivating heat-inactivated RNAP. GroEL673 can only protect RNAP, and this protecting ability is not stimulated by GroES. The mechanism by which the DnaJ and GrpE heat shock proteins contribute to DnaK's ability to reactivate heat-inactivated RNAP GroEL673 has also been investigated. We found that the DnaJ protein substantially reduces the levels of DnaK protein needed in this reactivation assay. However, the observed lag in reactivation is diminished only in the additional presence of the GrpE protein. Hence, DnaJ and GrpE are involved in both steps of this reactivation reaction (recognition of substrate and release of chaperone from the substrate-chaperone complex) while, in the case of the GroEL-dependent reaction, GroES is involved only during the release of chaperone from the substrate-chaperone complex.  相似文献   

13.
Chaperonins GroEL and GroES form two types of hetero-oligomers in vitro that can mediate the folding of proteins. Chemical cross-linking and electron microscopy showed that in the presence of adenosine triphosphate (ATP), two GroES7 rings can successively bind a single GroEL14 core oligomer. The symmetric GroEL14(GroES7)2 chaperonin, whose central cavity appears obstructed by two GroES7 rings, can nonetheless stably bind and assist the ATP-dependent refolding of RuBisCO enzyme. Thus, unfolded proteins first bind and possibly fold on the external envelope of the chaperonin hetero-oligomer.  相似文献   

14.
Pseudomonas aeruginosa has an anabolic (ArgF) and a catabolic (ArcB) ornithine carbamoyltransferase (OTCase). Despite extensive sequence similarities, these enzymes function unidirectionally in vivo. In the dodecameric catabolic OTCase, homotropic cooperativity for carbamoylphosphate strongly depresses the anabolic reaction; the residue Glu1O5 and the C-terminus are known to be essential for this cooperativity. When Glu1O5 and nine C-terminal amino acids of the catabolic OTCase were introduced, by in vitro genetic manipulation, into the closely related, trimeric, anabolic (ArgF) OTCase of Escherichia coli, the enzyme displayed Michaelis-Menten kinetics and no cooperativity was observed. This indicates that additional amino acid residues are required to produce homotropic cooperativity and a dodecameric assembly. To localize these residues, we constructed several hybrid enzymes by fusing, in vivo or in vitro, the E. coli argF gene to the P. aeruginosa arcB gene. A hybrid enzyme consisting of 101 N-terminal ArgF amino acids fused to 233 C-terminal ArcB residues and the reciprocal ArcB-ArgF hybrid were both trimers with little or no cooperativity. Replacing the seven N-terminal residues of the ArcB enzyme by the corresponding six residues of E. coli ArgF enzyme produced a dodecameric enzyme which showed a reduced affinity for carbamoylphosphate and an increase in homotropic cooperativity. Thus, the N-terminal amino acids of catabolic OTCase are important for interaction with carbamoylphosphate, but do not alone determine dodecameric assembly. Hybrid enzymes consisting of either 26 or 42 N-terminal ArgF amino acids and the corresponding C-terminal ArcB residues were both trimeric, yet they retained some homotropic cooperativity. Within the N-terminal ArcB region, a replacement of motif 28-33 by the corresponding ArgF segment destabilized the dodecameric structure and the enzyme existed in trimeric and dodecameric states, indicating that this region is important for dodecameric assembly. These findings were interpreted in the light of the three-dimensional structure of catabolic OTCase, which allows predictions about trimer-trimer interactions. Dodecameric assembly appears to require at least three regions: the N- and C-termini (which are close to each other in a monomer), residues 28-33 and residues 147-154. Dodecameric structure correlates with high carbamoylphosphate cooperativity and thermal stability, but some trimeric hybrid enzymes retain cooperativity, and the dodecameric Glu1O5-->Ala mutant gives hyperbolic carbamoylphosphate saturation, indicating that dodecameric structure is neither necessary nor sufficient to ensure cooperativity.  相似文献   

15.
A molecular mechanics study of polydehydroalanine [poly-(delta Ala)] is presented. For this purpose the AMBER 3a program has been used to perform the calculations. With exception of the point charges, the parameters for the terminal groups were taken from AMBER 3a libraries, whereas those for the delta Ala residue from Alagona et al. [J. Comp. Chem. (1991) Vol. 12, pp. 934-942]. Charges for the residue and terminal groups have been fitted from the MNDO electrostatic potential and scaled to achieve an ab initio 6-31G* quality. Calculations have been carried out using the continuous solvent approximation with three different dielectrics epsilon = 1, 1r, and 4r. The results show that, despite the preferred structure for the isolated residue is an extended conformation, a 310-helix is the preferred conformation in the solid state (epsilon = 1 and 1r), whereas a peculiar structure with psi = 0 degree is preferred with epsilon = 4r.  相似文献   

16.
As a basic principle, assisted protein folding by GroEL has been proposed to involve the disruption of misfolded protein structures through ATP hydrolysis and interaction with the cofactor GroES. Here, we describe chaperonin subreactions that prompt a re-examination of this view. We find that GroEL-bound substrate polypeptide can induce GroES cycling on and off GroEL in the presence of ADP. This mechanism promotes efficient folding of the model protein rhodanese, although at a slower rate than in the presence of ATP. Folding occurs when GroES displaces the bound protein into the sequestered volume of the GroEL cavity. Resulting native protein leaves GroEL upon GroES release. A single-ring variant of GroEL is also fully functional in supporting this reaction cycle. We conclude that neither the energy of ATP hydrolysis nor the allosteric coupling of the two GroEL rings is directly required for GroEL/GroES-mediated protein folding. The minimal mechanism of the reaction is the binding and release of GroES to a polypeptide-containing ring of GroEL, thereby closing and opening the GroEL folding cage. The role of ATP hydrolysis is mainly to induce conformational changes in GroEL that result in GroES cycling at a physiologically relevant rate.  相似文献   

17.
The cyanide-ligated form of the baker's yeast cytochrome c peroxidase mutant bearing the mutation Asn82-->Ala82 ([N82A]CcPCN) has been studied by proton NMR spectroscopy. This mutation alters an amino acid that forms a hydrogen bond to His52, the distal histidine residue that interacts in the heme pocket with heme-bound ligands. His52 is a residue critical to cytochrome c peroxidase's normal function. Proton hyperfine resonance assignments have been made for the cyanide-ligated form of the mutant by comparison with 1-D and NOESY spectra of the wild-type native enzyme. For [N82A]CcPCN, proton NMR spectra reveal two significant phenomena. First, similar to results published for the related mutant [N82D]CcPCN [Satterlee, J. D., et al. (1994) Eur. J. Biochem. 244, 81-87], for Ala82 mutation disrupts the hydrogen bond between His52 and the heme-ligated CN. Second, four of the 24 resolved hyperfine-shifted resonances are doubled in the mutant enzyme's proton spectrum, leading to the concept that the heme active site environment is dynamically microheterogeneous on a very localized scale. Two magnetically inequivalent enzyme forms are detected in a pure enzyme preparation. Varying temperature causes the two enzyme forms to interconvert. Magnetization transfer experiments further document this interconversion between enzyme forms and have been used to determine that the rate of interconversion is 250 (+/- 53) s-1. The equilibrium constant at 20 degrees C is 1.5. Equilibrium constants have been calculated at various temperatures between 5 and 29 degrees C leading to the following values: delta H = 60 kJ mol-1; delta S = 0.20 kJ K-1 mol-1.  相似文献   

18.
The chaperonins GroEL and GroES of Escherichia coli facilitate protein folding in an adenosine triphosphate (ATP)-dependent reaction cycle. The kinetic parameters for the formation and dissociation of GroEL-GroES complexes were analyzed by surface plasmon resonance. Association of GroES and subsequent ATP hydrolysis in the interacting GroEL toroid resulted in the formation of a stable GroEL:ADP:GroES complex. The complex dissociated as a result of ATP hydrolysis in the opposite GroEL toroid, without formation of a symmetrical GroEL:(GroES)2 intermediate. Dissociation was accelerated by the addition of unfolded polypeptide. Thus, the functional chaperonin unit is an asymmetrical GroEL:GroES complex, and substrate protein plays an active role in modulating the chaperonin reaction cycle.  相似文献   

19.
Here we report a method of immobilising the chaperonins GroEL and GroES to a glass matrix. The immobilised chaperone system has been used to successfully refold target proteins denatured by guanidine hydrochloride and produce substantially higher levels of active protein than occur on dilution into aqueous solution alone. The chaperone system has been shown to refold proteins from each of the three categories of GroEL substrate. The refolding of the enzyme glycerol dehydrogenase from Bacillus stearothermophilus shows a two-fold increase in activity in the presence of immobilised GroEL compared to that in free solution. The lactate dehydrogenase from B. stearothermophilus also shows a two-fold higher yield of activity in the presence of the immobilised GroEL and ATP. The presence of immobilised GroEL in the absence of ATP arrests the refolding of LDH. The enzyme citrate synthetase from porcine heart demonstrates a three-fold increase in activity when refolded in the presence of immobilised GroEL, ATP and free GroES. Similar results are obtained in the presence of free GroEL, immobilised GroES and ATP. The matrix-bound chaperone can be removed from the refolding mixture by centrifugation, producing a reusable system that can be easily isolated and purified from the refolded substrate.  相似文献   

20.
Regulatory light chain (RLC) mutants, RLC-C18 and RLC-C165, containing a single cysteine at positions 18 and 165 near the N and C terminus, respectively, were each labeled with benzophenone 4-iodoacetamide and exchanged into myosin in their phosphorylated or unphosphorylated forms and then photolyzed. SDS-PAGE showed that, for RLC-C18, the intrachain photo-cross-linking in myosin was inhibited by phosphorylation. For myosin containing RLC-C165, the yield of one intrachain cross-linked band decreased significantly whereas the other was unaffected by phosphorylation. Peptide mapping in conjunction with mass spectrometry showed that Cys165 was cross-linked to site(s) within Ala17-Lys34 independent of the phosphorylation of Ser19. This clearly demonstrates that the proximity between the N- and C-terminal regions of RLC is not affected by phosphorylation. In addition, Cys165 could also be cross-linked to the region of Phe133-Arg143; however, this type of cross-linking was inhibited in the phosphorylated state. For RLC-C18, the cross-linking took place with the region of Glu124-Arg132 or Phe133-Arg143, also only in the unphosphorylated state. Thus, phosphorylation changes the spatial relationship between the region of Glu124-Arg143 and Cys18 and Cys165. In scallop myosin, the region corresponding to Glu124-Arg143 is located at the interfaces between RLC and the essential light chain as well as the heavy chain [Xie, X. , et al. (1994) Nature 368, 306-312]. In light of that work, our results suggest that the region of Glu124-Arg143 is involved in the phosphorylation-dependent signaling and the change in its spatial relationship with respect to the N and C termini of RLC may underlie the activation of the smooth muscle myosin.  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