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 0 毫秒
1.
J Mackall  C B Klee 《Biochemistry》1991,30(29):7242-7247
The rate of proteolysis of trypsin-sensitive bonds was used to examine the nature of the structural changes accompanying Ca2+ and Mg2+ binding to calmodulin. In the Ca(2+)-free form, the rates of proteolysis at Arg-106 and Arg-37 are rapid (greater than 300 and 28 nmol min-1 mL-1, respectively), the bonds at Arg-74, Lys-75, and Lys-77, in the central helix, are cleaved more slowly (10 nmol min-1 mL-1), and a lag in the cleavage at the remaining bonds (Lys-13, Lys-30, Arg-86, Arg-90, and Arg-126) suggests that they are not cleaved in the native protein. High concentrations of Ca2+, but not Mg2+, almost completely abolish proteolysis at Arg-106 and drastically reduce the rate of cleavage at Arg-37. Both Ca2+ and Mg2+ exert a moderate protective effect on the proteolysis of the central helix. These results suggest that the F-helix of domains III and, to a lesser extent, the F-helix of domain I are somewhat flexible in the Ca(2+)-free form and are stabilized by Ca2+. Whereas full occupancy of the four Ca(2+)-binding sites produces little change in the susceptibility of the central helix to proteolytic attack, binding of two Ca2+ produces a 10-fold enhancement of the rate of proteolysis in this part of the molecule. We propose that at intermediate Ca2+ levels the flexibility of the central helix of calmodulin is greatly increased, resulting in the transient formation of intermediates which have not been detected by spectroscopic techniques but are trapped by the irreversible action of trypsin.  相似文献   

2.
Bending of the calmodulin central helix: a theoretical study.   总被引:3,自引:2,他引:3  
The crystal structure of calcium-calmodulin (CaM) reveals a protein with a typical dumbbell structure. Various spectroscopic studies have suggested that the central linker region of CaM, which is alpha-helical in the crystal structure, is flexible in solution. In particular, NMR studies have indicated the presence of a flexible backbone between residues Lys 77 and Asp 80. This flexibility is related directly to the function of the protein because it enables the N- and C-terminal domains of the protein to move toward each other and bind to the CaM-binding domain of a target protein. We have investigated the flexibility of the CaM central helix by a variety of computational techniques: molecular dynamics (MD) simulations, normal mode analysis (NMA), and essential dynamics (ED) analysis. Our MD results reproduce the experimentally determined location of the bend in a simulation of only the CaM central helix, indicating that the bending point is an intrinsic property of the alpha-helix, for which the remainder of the protein is not important. Interestingly, the modes found by the ED analysis of the MD trajectory are very similar to the lowest frequency modes from the NM analysis and to modes found by an ED analysis of different structures in a set of NMR structures. Electrostatic interactions involving residues Arg 74 and Asp 80 seem to be important for these bending motions and unfolding, which is in line with pH-dependent NMR and CD studies.  相似文献   

3.
The central helix of calmodulin functions as a flexible tether   总被引:13,自引:0,他引:13  
Using site-directed mutagenesis we have created an altered calmodulin in which Gln-3 and Thr-146 have both been replaced by cysteines. We have reacted this protein with the bifunctional reagent, bismaleimidohexane, forming an intramolecular cross-link between the two cysteines. In the crystal structure of native calmodulin alpha-carbons at positions 3 and 146 are 37 A apart. In the bismaleimidohexane cross-linked protein these atoms can be no more than 19 A apart, and model building studies indicate that there is probably a bend in the central helix of calmodulin. A second modified calmodulin was generated by cleaving the central helix of the cross-linked protein at Lys-77 with trypsin. In this molecule, the two lobes of calmodulin are joined solely by the bismaleimidohexane cross-link, which bridges Cys-3 and Cys-146. Vm and Kact values for activation of myosin light chain kinase activity by the cross-linked and cross-linked/trypsinized proteins are not significantly different from those for the control protein. This result indicates that one role for the central helix may be to serve as a flexible tether between the calmodulin lobes. This is consistent with a model calmodulin-enzyme complex in which the central helix is bent, and the two lobes exert a concerted effect. A detailed model of this type has been proposed for the calmodulin-myosin light chain kinase complex (Persechini, A. and Kretsinger, R.H. (1988) J. Cardiovasc. Pharmacol., in press).  相似文献   

4.
Calmodulin is phosphorylated by casein kinase II on Thr-79, Ser-81, Ser-101 and Thr-117. To determine the consensus sequences for casein kinase II in intact calmodulin, we examined casein kinase II-mediated phosphorylation of engineered calmodulins with 1-4 deletions in the central helical region (positions 81-84). Total casein kinase II-catalyzed phosphate incorporation into all deleted calmodulins was similar to control calmodulin. Neither CaM delta 84 (Glu-84 deleted) nor CaM delta 81-84 (Ser-81 to Glu-84 deleted) has phosphate incorporated into Thr-79 or Ser-81, but both exhibit increased phosphorylation of residues Ser-101 and Thr-117. These data suggest that phosphoserine in the +2 position may be a specificity determinant for casein kinase II in intact proteins and/or secondary structures are important in substrate recognition by casein kinase II.  相似文献   

5.
H Sun  D Yin  T C Squier 《Biochemistry》1999,38(38):12266-12279
We have used fluorescence spectroscopy to investigate the average structure and extent of conformational heterogeneity associated with the central helix in calmodulin (CaM), a sequence that contributes to calcium binding sites 2 and 3 and connects the amino- and carboxyl-terminal globular domains. Using site-directed mutagenesis, a double mutant was constructed involving conservative substitution of Tyr(99) --> Trp(99) and Leu(69) --> Cys(69) with no significant effect on the secondary structure of CaM. These mutation sites are at opposite ends of the central helix. Trp(99) acts as a fluorescence resonance energy transfer (FRET) donor in distance measurements of the conformation of the central helix. Cys(69) provides a reactive group for the covalent attachment of the FRET acceptor 5-((((2-iodoacetyl)amino)ethyl)amino)naphthalene-1-sulfonic acid (IAEDANS). AEDANS-modified CaM fully activates the plasma membrane (PM) Ca-ATPase, indicating that the native structure is retained following site-directed mutagenesis and chemical modification. We find that the average spatial separation between Trp(99) and AEDANS covalently bound to Cys(69) decreases by approximately 7 +/- 2 A upon calcium binding. However, irrespective of calcium binding, there is little change in the conformational heterogeneity associated with the central helix under physiologically relevant conditions (i.e., pH 7.5, 0.1 M KCl). These results indicate that calcium activation alters the spatial arrangement of the opposing globular domains between two defined conformations. In contrast, under conditions of low ionic strength or pH the structure of CaM is altered and the conformational heterogeneity of the central helix is decreased upon calcium activation. These results suggest the presence of important ionizable groups that affect the structure of the central helix, which may play an important role in mediating the ability of CaM to rapidly bind and activate target proteins.  相似文献   

6.
Using site-directed mutagenesis we have expressed in Escherichia coli three engineered calmodulins (CaM) containing deletions in the solvent-exposed region of the central helix. These are CaM delta 84, Glu-84 removed; CaM delta 83-84, Glu-83 and Glu-84 removed; and CaM delta 81-84, Ser-81 through Glu-84 removed. The abilities of these proteins to activate skeletal muscle myosin light chain kinase, plant NAD kinase, and bovine brain calcineurin activities were determined, as were their abilities to bind a synthetic peptide based on the calmodulin-binding domain of skeletal muscle myosin light chain kinase. Similar results were obtained with all three deletion proteins. Vm values for enzymes activated by the deletion proteins are all within 10-20% of those values obtained with bacterial control calmodulin. Relative to bacterial control values, changes in Kact or Kd values associated with the deletions are all less than an order of magnitude: Kact values for NAD kinase and myosin light chain kinase are increased 5-7-fold, Kd values for binding of the synthetic peptide are increased 4-7-fold, and Kact values for calcineurin are increased only 1-3-fold. In assays of NAD kinase and myosin light chain kinase activation some differences between bovine calmodulin and bacterial control calmodulin were observed. With NAD kinase, Kact values for the bacterial control protein are increased 4-fold relative to values for bovine calmodulin, and Vm values are increased by 50%; with myosin light chain kinase, Kact values are increased 2-fold and Vm values are decreased 10-15% relative to those values obtained with bovine calmodulin. These differences between bacterial control and bovine calmodulins probably can be attributed to known differences in postranslational processing of calmodulin in bacterial and eucaryotic cells. No differences between bovine and control calmodulins were observed in assays of calcineurin activation or peptide binding. Our observations indicate that contacts with the deleted residues, Ser-81 through Glu-84, are not critical in the calmodulin-target complexes we have evaluated. Formation of these calmodulin-target complexes also does not appear to be greatly affected by the global alterations in the structure of calmodulin that are associated with the deletions. In models in which the central helix is maintained in the altered calmodulins, each deleted residue causes the two lobes of calmodulin to be twisted 100 degrees relative to one another and brought 1.5 A closer together.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
Calmodulin (CaM) and troponin C (TnC) are the most similar members of EF-hand family and show few differences in the primary structure. Here, we use mutants of troponin that mimic calmodulin and changes in temperature to investigate the factors that determine their specificity as regulatory proteins. Using a double mutant of troponin that resembles calmodulin in lacking both the N-terminal helix and KGK(91-93) we observe a small difference from troponin in binding to the erythrocyte Ca(2+)-ATPase, and an improvement in enzyme activation. A triple mutant, where in addition, the residues 88-90 are replaced with the corresponding sequence from calmodulin is equivalent to calmodulin in maximal activation, and it restores protein ability to increase Ca(2+) affinity for the enzyme. However, this mutant also binds less tightly (1/100) than calmodulin. Remarkably, a decrease in temperature has a more marked effect in protein binding than either mutation, reducing the difference in affinities to 18-fold, but without any improvement in their ability to increase Ca(2+) affinity for the enzyme. Spectroscopic analysis of hydrophobic domain exposure in EF-hand proteins was carried out using 8-anilino-1-naphthalenesulfonic acid (ANS). The probe shows a much higher fluorescence when bound to the complex Ca(4)-calmodulin than to Ca(4)-troponin. Decreasing the temperature exposes additional hydrophobic regions of troponin. Changing the Mg(2+) concentration does not affect their bindings to the enzyme. It is suggested that the requirements for troponin to mimic calmodulin in binding to the target enzyme, and those for activating it, are met by different regions of the protein.  相似文献   

8.
Functional expression of chicken calmodulin in yeast   总被引:10,自引:0,他引:10  
The coding region of a chicken calmodulin cDNA was fused to a galactose-inducible GAL1 promoter, and an expression system was constructed in the yeast Saccharomyces cerevisiae. Expression of calmodulin was demonstrated by purifying the heterologously expressed protein and analyzing its biochemical properties. When the expression plasmid was introduced into a calmodulin gene (cmd1)-disrupted strain of yeast, the cells grew in galactose medium, showing that chicken calmodulin could complement the lesion of yeast calmodulin functionally. Repression of chicken calmodulin in the (cmd1)-disrupted strain caused cell cycle arrest with a G2/M nucleus, as observed previously with a conditional-lethal mutant of yeast calmodulin. These results suggest that the essential function of calmodulin for cell proliferation is conserved in cells ranging from yeast to vertebrate cells.  相似文献   

9.
G Barbato  M Ikura  L E Kay  R W Pastor  A Bax 《Biochemistry》1992,31(23):5269-5278
The backbone dynamics of Ca(2+)-saturated recombinant Drosophila calmodulin has been studied by 15N longitudinal and transverse relaxation experiments, combined with 15N(1H) NOE measurements. Results indicate a high degree of mobility near the middle of the central helix of calmodulin, from residue K77 through S81, with order parameters (S2) in the 0.5-0.6 range. The anisotropy observed in the motion of the two globular calmodulin domains is much smaller than expected on the basis of hydrodynamic calculations for a rigid dumbbell type structure. This indicates that, for the purposes of 15N relaxation, the tumbling of the N-terminal (L4-K77) and C-terminal (E82-S147) lobes of calmodulin is effectively independent. A slightly shorter motional correlation time (tau c approximately 6.3 ns) is obtained for the C-terminal domain compared to the N-terminal domain (tau c approximately 7.1 ns), in agreement with the smaller size of the C-terminal domain. A high degree of mobility, with order parameters of approximately 0.5, is also observed in the loop that connects the first with the second EF-hand type calcium binding domain and in the loop connecting the third and fourth calcium binding domain.  相似文献   

10.
Alteration of residues 82-84 in the alpha-helix that links the two halves of calmodulin results in a differential effect on activator activity. Previous studies (Lukas, T. J., Burgess, W. H., Prendergast, F. G., Lau, W., and Watterson, D. M. (1986) Biochemistry 25, 1458-1464) indicated the importance of positive charge clusters in the calmodulin-binding protein, myosin light chain kinase. This suggested the possible importance of complementary negative charge clusters in calmodulin. By using an efficient cassette mutagenesis approach and a synthetic calmodulin gene (Roberts, D. M., Crea, R., Malecha, M., Alvarado-Urbina, G., Chiarello, R. H., and Watterson, D. M. (1985) Biochemistry 24, 5090-5098), this possibility was directly addressed by engineering a new calmodulin, VU-8 calmodulin, in which the glutamate cluster at residues 82-84 in the synthetic gene product (VU-1 calmodulin) was replaced by three lysines. VU-8 calmodulin activated phosphodiesterase to the same maximal extent as VU-1 calmodulin, although there was an alteration in the concentration of calmodulin required for half-maximal stimulation. In contrast, myosin light chain kinase was activated to only 30% of maximal activity and NAD kinase was not activated. These results provide insight into the functional role of the unusual central helix structure found in the calmodulin family of proteins and indicate that different, although possibly overlapping, chemical complementarities are employed in the interaction between calmodulin and its various physiological targets.  相似文献   

11.
Neuromodulin (also designated P-57, GAP-43, B-50) is a major presynaptic substrate for protein kinase C. Phosphorylation of neuromodulin decreases its affinity for calmodulin, suggesting that neuromodulin may function to bind and concentrate calmodulin at specific sites within neurons, releasing calmodulin locally in response to phosphorylation by protein kinase C (Alexander, K. A., Cimler, B. M., Meier, K. E., and Storm, D. R. (1987) J. Biol. Chem. 262, 6108-6113). In the present study, we have constructed and characterized several mutant neuromodulins to demonstrate that the amino acid sequence 39-56 is required for calmodulin binding, and that this domain contains the sole in vitro protein kinase C phosphorylation site at serine 41. We also demonstrate that the adjacent phenylalanine 42, interacts hydrophobically with calmodulin. These hydrophobic interactions may be disrupted by the introduction of negative charge at serine 41, and thereby regulate the neuromodulin/calmodulin binding interactions. The sensitivity of the neuromodulin/calmodulin binding interaction to negative charge at serine 41 was determined by substitution of serine 41 with an aspartate or an asparagine residue. The asparagine mutant retained its affinity for calmodulin-Sepharose while the aspartate mutant did not adsorb to calmodulin-Sepharose. We conclude that protein kinase C phosphorylation of neuromodulin abolishes calmodulin binding by introducing negative charges within the calmodulin binding domain at a position adjacent to the phenylalanine.  相似文献   

12.
The second half of the photocycle of the light-driven proton pump bacteriorhodopsin includes proton transfers between D96 and the retinal Schiff base (the M to N reaction) and between the cytoplasmic surface and D96 (decay of the N intermediate). The inhibitory effects of decreased water activity and increased hydrostatic pressure have suggested that a conformational change resulting in greater hydration of the cytoplasmic region is required for proton transfer from D96 to the Schiff base, and have raised the possibility that the reversal of this process might be required for the subsequent reprotonation of D96 from the cytoplasmic surface. Tilt of the cytoplasmic end of helix F has been suggested by electron diffraction of the M intermediate. Introduction of bulky groups, such as various maleimide labels, to engineered cysteines at the cytoplasmic ends of helices A, B, C, E, and G produce only minor perturbation of the decays of M and N, but major changes in these reactions when the label is linked to helix F. In these samples the reprotonation of the Schiff base is accelerated and the reprotonation of D96 is strongly retarded. Cross-linking with benzophenone introduced at this location, but not at the others, causes the opposite change: the reprotonation of the Schiff base is greatly slowed while the reprotonation of D96 is accelerated. We conclude that, consistent with the structure from diffraction, the proton transfers in the second half of the photocycle are facilitated by motion of the cytoplasmic end of helix F, first away from the center of the protein and then back.  相似文献   

13.
Molecular dynamics studies of the N-domain (amino acids 1-77; CaM(1-77)) of Ca2+-loaded calmodulin (CaM) show that a solvent exposed hydrophobic cleft in the crystal structure of CaM exhibits transitions from an exposed (open) to a buried (closed) state over a time scale of nanoseconds. As a consequence of burying the hydrophobic cleft, the R(g) of the protein is reduced by 1.5 A. Based on this prediction, x-ray scattering experiments were conducted on this domain over a range of concentrations. Models built from the scattering data show that the R(g) and general shape is consistent with the simulation studies of CaM(1-77). Based on these observations we postulate a model in which the conformation of CaM fluctuates between two different states that expose and bury this hydrophobic cleft. In aqueous solution the closed state dominates the population, while in the presence of peptides, the open state dominates. This inherent flexibility of CaM may be the key to its versatility in recognizing structurally distinct peptide sequences. This model conflicts with the currently accepted hypothesis based on observations in the crystal structure, where upon Ca2+ binding the hydrophobic cleft is exposed to solvent. We postulate that crystal packing forces stabilize the protein conformation toward the open configuration.  相似文献   

14.
Two mutant forms of calmodulin were examined by small-angle X-ray scattering in solution and compared with the wild-type protein. Each mutant has deletions in the linker region of the central helix: one lacks residues Glu-83 and Glu-84 (Des2) and the other lacks residues Ser-81 through Glu-84 (Des4). The deletions change both the radii of gyration and the maximum dimensions of the molecules. In the presence of Ca2+, the observed radii of gyration are 22.4 A for wild-type bacterially expressed calmodulin, 19.5 A for Des2 calmodulin, and 20.3 A for Des4 calmodulin. A reduction in the radius of gyration by 1-2 A on removal of calcium, previously observed in the native protein, was also found in the wild type and the Des4 mutant; however, no significant size change was observed in the Des2 mutant. The large calcium-dependent conformational change in calmodulin induced by the binding of melittin [Kataoka, M., Head, J.F., Seaton, B.A., & Engelman, D.M. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 6944-6948] was observed in all the bacterially expressed proteins. Each protein appears to undergo a transition from a dumbbell shape to a more globular conformation on binding melittin in the presence of calcium, although quantitatively the changes in the wild-type and Des4 proteins greatly exceed those in Des2. Modeling shows the central linker region of the molecule. Thus, the structure of the linker region is stable enough to maintain the average orientation and separation of the lobes yet flexible enough to permit the lobes to approach each other upon binding a peptide.  相似文献   

15.
M Ikura  L E Kay  M Krinks  A Bax 《Biochemistry》1991,30(22):5498-5504
Heteronuclear 3D and 4D NMR experiments have been used to obtain 1H, 13C, and 15N backbone chemical shift assignments in Ca(2+)-loaded calmodulin complexed with a 26-residue synthetic peptide (M13) corresponding to the calmodulin-binding domain (residues 577-602) of rabbit skeletal muscle myosin light-chain kinase. Comparison of the chemical shift values with those observed in peptide-free calmodulin [Ikura, M., Kay, L. E., & Bax, A. (1990) Biochemistry 29, 4659-4667] shows that binding of M13 peptide induces substantial chemical shift changes that are not localized in one particular region of the protein. The largest changes are found in the first helix of the Ca(2+)-binding site I (E11-E14), the N-terminal portion of the central helix (M72-D78), and the second helix of the Ca(2+)-binding site IV (F141-M145). Analysis of backbone NOE connectivities indicates a change from alpha-helical to an extended conformation for residues 75-77 upon complexation with M13. This conformational change is supported by upfield changes in the C alpha and carbonyl chemical shifts of these residues relative to M13-free calmodulin and by hydrogen-exchange experiments that indicate that the amide protons of residues 75-82 are in fast exchange (kexch greater than 10 s-1 at pH 7, 35 degrees C) with the solvent. No changes in secondary structure are observed for the first helix of site I or the C-terminal helix of site IV. Upon complexation with M13, a significant decrease in the amide exchange rate is observed for residues T110, L112, G113, and E114 at the end of the second helix of site III.  相似文献   

16.
17.
Calcium-dependent changes in the internal dynamics and average structures of the opposing globular domains of calmodulin (CaM), as well as their relative spatial arrangement, contribute to the productive association between CaM and a range of different target proteins, affecting their functional activation. To identify dynamic structural changes involving individual alpha-helical elements and their modulation by calcium activation, we have used site-directed mutagenesis to engineer a tetracysteine binding motif within helix A near the amino terminus of calmodulin (CaM), permitting the selective and rigid attachment of the fluorescent probe 4',5'-bis(1,3,2-dithioarsolan-2-yl)fluorescein (FlAsH) with full retention of function. The rigid tetracoordinate linkage of FlAsH to CaM, in conjunction with frequency domain fluorescence anisotropy measurements, allows assessment of dynamic changes associated with calcium activation without interference from independent probe motion. Taking advantage of the large fluorescence enhancement associated with binding of FlAsH to CaM, we determined rates of binding of FlAsH to apo-CaM and calcium-activated CaM to be 2800 +/- 80 and 310 +/- 10 M(-)(1) s(-)(1), respectively. There is no difference in the solvent accessibility of the bound FlAsH irrespective of calcium binding to CaM. Thus, given that FlAsH selectively labels disordered structures, the large difference in rates of FlAsH binding indicates that calcium binding stabilizes helix A. Frequency domain anisotropy measurements of bound FlAsH indicate that prior to calcium activation, helix A undergoes large amplitude nanosecond motions. Following calcium activation, helix A becomes immobile, and structurally coupled to the overall rotation of CaM. We discuss these results in the context of a model that suggests stabilization of helix A relative to other domain elements in the CaM structure is critical to defining high-affinity binding clefts, and in promoting specific and ordered binding of the opposing lobes of CaM to target proteins.  相似文献   

18.
The structural significance of C-terminal amphiphilic alpha-helix of human interleukin-2 has been investigated using principles of protein design. Employing disulfide-mediated semi-synthesis, several multiple residue substitution patterns were studied in order to provide rapid insight into the most appropriate features to incorporate into fully recombinant proteins. Substitutions directed toward both stabilization and destabilization of the helix resulted in proteins with modulated bioactivity. Circular dichroism verified the conformational integrity and thermal stability of the derivatives. The biologic characteristics of each derivative were evaluated in the standard murine CTLL-2 assay and compared to activities exhibited in both human T-cell bioactivity and binding assay. A strategy for the design of protein ligand agonists and antagonists without knowledge of receptor contact residues is discussed.  相似文献   

19.
20.
This work considers the response to simulated synaptic inputs of an excitable membrane model. The model is essentially of the Hodgkin-Huxley type, but contains an A-current in addition to sodium and delayed-rectifier potassium channels. The results were compared with previous simulations in which the stimulus was an injected current. These two types of stimuli give somewhat different results because synaptic stimuli directly change the membrane resistance, whereas injected current does not. The results of synaptic stimulation were similar to injected current in that very low frequencies of action potentials were elicited only where the stimulus was slightly above threshold. For most of the range of synaptic inputs that produced oscillatory behavior, the A-current had little effect on oscillation frequency. With synaptic stimuli as with injected current, the model membrane's spiking behavior does not begin immediately when an excitatory stimulus is imposed on a quiescent state. The delay before spiking is closely related to the inactivation time of the A-current. The synaptic results were different from the injected current results in that when substantial inhibition was present, the ability to produce very-low-frequency spiking was absent, even just above the excitatory threshold. The higher the degree of inhibition, the narrower the range of spike frequencies that could be elicited by excitation. At very high inhibition, no degree of excitation could elicit spiking.  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